Next Article in Journal
Prognostic Biomarkers of Cell Proliferation in Colorectal Cancer (CRC): From Immunohistochemistry to Molecular Biology Techniques
Next Article in Special Issue
A Narrative Review on CD44’s Role in Glioblastoma Invasion, Proliferation, and Tumor Recurrence
Previous Article in Journal
Treatment of Colorectal Cancer in Certified Centers: Results of a Large German Registry Study Focusing on Long-Term Survival
Previous Article in Special Issue
Serum Paraprotein Is Associated with Adverse Prognostic Factors and Outcome, across Different Subtypes of Mature B-Cell Malignancies—A Systematic Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Clinical Validation of Modulated Electro-Hyperthermia (mEHT)

1
Department of Radiation Oncology, Jeonbuk National University Medical School, Jeonju 54907, Republic of Korea
2
Research Institute of Clinical Medicine of Jeonbuk National University-Biomedical Research Institute of Jeonbuk National University Hospital, Jeonju 54907, Republic of Korea
3
Division of Oncology, Department of Internal Medicine and Oncology, Semmelweis University, H-1083 Budapest, Hungary
4
Faculty of Information Technology and Bionics, Pázmány Péter Catholic University, H-1083 Budapest, Hungary
*
Author to whom correspondence should be addressed.
Cancers 2023, 15(18), 4569; https://doi.org/10.3390/cancers15184569
Submission received: 30 May 2023 / Revised: 30 August 2023 / Accepted: 4 September 2023 / Published: 15 September 2023
(This article belongs to the Special Issue Advanced Research in Oncology in 2023)

Abstract

:

Simple Summary

Modulated electro-hyperthermia (mEHT) is a heating therapy that uses synergized thermal and nonthermal effects to heat and destroy malignant cells selectively without damaging healthy cells. This article presents the clinical validation of mEHT. The therapy is dominantly applied for such advanced malignancies when the conventional oncotherapies fail to apply. Survival results of mEHT were collected and compared with other methods. The results demonstrate the superiority of the mEHT method.

Abstract

The mEHT method uses tissues’ thermal and bioelectromagnetic heterogeneity for the selective mechanisms. The success of the therapy for advanced, relapsed, and metastatic aggressive tumors can only be demonstrated by measuring survival time and quality of life (QoL). The complication is that mEHT-treated patients cannot be curatively treated any longer with “gold standards”, where the permanent progression of the disease, the refractory, relapsing situation, the organ failure, the worsening of blood counts, etc., block them. Collecting a cohort of these patients is frequently impossible. Only an intent-to-treat (ITT) patient group was available. Due to the above limitations, many studies have single-arm data collection. The Phase III trial of advanced cervix tumors subgrouping of HIV-negative and -positive patients showed the stable efficacy of mEHT in all patients’ subgroups. The single-arm represents lower-level evidence, which can be improved by comparing the survival data of various studies from different institutes. The Kaplan–Meier probability comparison had no significant differences, so pooled data were compared to other methods. Following this approach, we demonstrate the feasibility and superiority of mEHT in the cases of glioblastoma multiform, pancreas carcinomas, lung tumors, and colorectal tumors.

Graphical Abstract

1. Introduction

The history of clinical hyperthermia can be traced back to the past. Ancient Greek medicine already used the method to treat oncological cases without detailed knowledge about the physiological feedback of the human body. Knowing the nature of the febrile condition, which in many cases was the guarantee of recovery, it was believed that proper heating would solve most medical problems [1]. They trusted that the body’s reactions would lead to the heating effect, inducing self-healing reactions of the organism. They essentially took advantage of the system’s striving for homeostatic balance, which is facilitated by heat stimuli. This principle is still appropriate in modern medicine, considering the limits of complex natural regulation.
Homeostatic surveillance controls the system’s stability and adaptability. The local or systemic heating interrupts the regulatory processes of stability, reacting with non-linear physiological responses to correct the inconsistency [2]. Theoretical biology often ignores this complex control facing a tragicomedy challenge [3]. The homeostatic control tries to re-establish the unheated conditions by non-linear feedback, increasing the cooling blood flow (BF) [4,5] as an effective heat exchanger. This complex dynamic behavior guarantees the robust stability of health conditions, so the reactive BF challenges the heating processes. Homeostatic control tries to restore healthy regulation by increasing blood flow and vasodilatation. However, the risk of invasion of tumor cells is enhanced by these corrective effects and may promote malignant dissemination.
Modulated electro-hyperthermia (mEHT) aims at a harmonic solution to this contradiction [6]. It applies electromagnetic interactions to deliver energy to the tumor. The energy is realized in the synergy of two basic effects:
  • Thermal effects occur in the form of heat and temperature increase. Thermal effects are mostly unselective; the heat spreads all over the volume seeking thermal equilibrium. The temperature characterizes the homogeneous distribution as average energy of the heat-absorbers.
  • Nonthermal processes are electron excitations, generating chemical reactions. The nonthermal impact may change the intercellular membrane, and intracellular processes select them by the dielectric and conductive heterogeneity of the target.
The mEHT applies a precise, personalized theranostic selection and treatment of malignancy, supporting natural homeostatic processes such as apoptosis, immune reactions, conditional effects, etc. [7]. The selection of the malignant cells uses the microscopic natural heterogeneities of the tumor. The applied electric field has different interactions with the cancerous and healthy cells in four basic characteristics, Figure 1, discussed in multiple publications [8,9,10,11,12]:
  • Due to the intensive metabolism of the malignant cells, the ionic species of the nutrients and waste molecules (such as lactate) have high concentrations in the tumor microenvironment (TME), and together with the extended volume of the extracellular matrix (ECM), create a significantly higher electric conductivity of the microenvironment of malignant cells and the entire tumor [13,14]. This conduction difference drives the RF current to the area [15].
  • The malignant cells break their networking connections (e.g., adherent connections and junctions [16]), and became autonomic. This cellular individualism makes the tumor microenvironment different, causing a higher dielectric permittivity of the tumor microenvironment (TME) than it was in the networking conditions [17,18]. The high dielectric permittivity favors conducting the radiofrequency (RF), making an additional selective factor for tumor cells.
  • The broken connections leave numerous transmembrane proteins on the membrane of the malignant cells. These membrane-embedded proteins and their lipid-enriched clusters (membrane rafts) have significantly higher energy absorption from the RF current than their surrounding lipid layer [19]. This makes these proteins particularly heatable and chemically excitable.
  • The malignancy had lost its healthy homeostatic control, and so it has locally modified physiologic regulations [20]. The arising structural and pathological modifications appear as an additional selectivity factor.
The auto-selection is theranostic, finding and treating the malignancy in macro- and micro-regions. The theranostic impact has special enhancing factors:
  • The particular energy intake of membrane rafts of malignant cells selectively heats them, working similar to natural absorbing nanoparticles [21]. This makes effective micro- and macro-localization of the heat effect [22].
  • The frequency dispersion has an optimal range of RF application for the above selection. However, the requested optimal frequency range of the membrane energy absorption/excitation and the driving of the molecular changes during the excitation need different frequencies which must be coordinated. The selection absorption optimum is near 10 MHz [23], while the desired molecular changes happen with a frequency less than 10 kHz. This 1/1000 ratio may be solved by modulation. The carrier is the approved medical frequency 13.56 MHz, and the modulation is a spectrum in the 10 Hz–10 kHz region [24]. The modulation spectrum is the physiologic noise of healthy homeostasis (its power density depends on the reciprocal value of the frequency) [25], and so forces the homeostatic control.
  • The applied modulated RF current kills the malignant cells in an apoptotic way, producing a damage-associated molecular pattern (DAMP) [26], realizing an immunogenic cell death (ICD) [27]. The ICD secrete calreticulin (CRT) [28] and heat-shock protein (HSP) on the malignant membrane [29] and attract the natural killer cells [30]; this is proven with mEHT too [31]. The ICD liberates the high mobility group box 1 (HMGB1) molecules [32] together with HSP70, HSP90 [33], and ATP [34] into the ECM. The membrane HSP-s activate the natural killer cells, and the other DAMP molecules maturate the dendritic cells, producing antigen presentation which creates immune reactions. The rising tumor-specific killer and helper T-cells activate antitumoral processes all over the body, acting on distant micro- and macro-metastases (abscopal effect) [35].
The preclinically proven selective mEHT processes [36] have numerous clinical studies [37]. The clinical efficacy of these trials is focused on patient-centered values: survival time and quality of life.
A broad spectrum of cancers shows the practical applicability of mEHT in human oncotherapies [37], which validates the preclinical molecular results [36] and supports the method’s feasibility [9]. The applied low incident energy is enough to detect, select, and treat the tumors in a theranostic way, irrespective of their form and size, and reach significant clinical achievements through the selected cellular heating [38].
The mEHT application in human clinical practice showed typical thermally enhanced BF measured in cervical carcinoma [39]. The produced mild 38.5℃ temperature in cervical cancer looks optimal for the complementary treatments because it provides enough blood support but is not too high to increase the dissemination of the malignant cells. Furthermore, the synergy of the thermal and nonthermal processes improves the pharmacokinetics measured in healthy volunteers [40,41].

2. Clinical Validation

2.1. Crossroads of Clinical Applications

The general cancer curative strategy is based on the direct distortion of the detected cancer cells by surgery, chemo, and radiotherapy. The distortion strategy offers a proper solution when the method is effective and selective enough. Selectivity means that it destroys all cancer cells, does not dangerously affect its healthy neighborhood, and does not escalate to paralyze essential body functions. The accuracy of the detection of malignant cells, including the disseminated ones, limits the destruction strategy because the remaining cancer cells may redevelop the disease and even could build resistance to the applied chemo and radio methods. These challenges create a new approach, using the system’s defense and protective procedures against malignancy and mobilizing the body’s immune system through immuno-oncology.

2.1.1. Change of Paradigm

We are in a war against cancer. The old military rule requests the attack of the enemy’s weakest point, but the conventional oncotherapies, including hyperthermia, attack the strongest malignancy side: proliferation. Change is necessary. Attack the weakest side: the missing networking, and consequently, the cancer is out from the overall regulation of the system. The cancer cells lost their collective connections and became autonomic.
Furthermore, the tumor-oriented curative approach has to be changed to the patient-oriented strategy to cure the patient, taking care of their complex health issues. Therapy must concentrate on human complexity and turn the product-oriented focus to the process-oriented one. This approach means that instead of concentrating on some molecular products, such as heat shock proteins, angiogenesis blockers, proliferation blockers, ionic blockers, blood-flow blockers, cell-poisoning, etc., we must focus on the dynamism of the cancer evolution process considering the immune effects, the physiologic feedback, and the overall homeostatic surveillance.

2.1.2. Clinical Challenges

Unlike conventional hyperthermia, the mEHT treatment is recommended for:
  • Patients who cannot receive either surgery, chemo, or radiotherapy (conventional gold standards) according to various contraindicated aspects such as:
    • They have a comorbidity that contraindicates the conventional oncotherapy procedures.
    • There is no effective conventional procedure for the given tumor.
  • Conventional curative procedures are no longer available and usually get a palliative setting only.
    • Patients with relapsed locally far-advanced tumors and no alternative standard curative therapy exists.
    • Conventional therapy cannot be continued due to organ failure or low blood count.
    • Despite standard treatments, patients show intense progression, relapse, and broad malignant dissemination.
  • Severe metastatic activity does not allow conventional treatment, salvage, or terminal state.
Due to the above challenges, most of the clinical trials conducted with mEHT are single-arm prospective or retrospective observational studies, meaning that it does not differ from other studies of medical devices [42].
The frequently applied single-arm studies have ethical and statistical reasons. Patients in this stage have no other treatments possible, and a cohort reference frequently does not exist for complete study statistics.

2.1.3. Study Challenges

The evidence from single-arm observational studies is usually less convincing than that from randomized double-arm studies due to the huge variation of personal situations for patients with severe advanced stages of the disease. The personalized protocol cannot be rigidly fixed in preliminary planning. Having personal variations of the protocol makes the single-arm prospective trial difficult because the patient cohort and its protocol are not homogeneous. In these cases, only an observational study is available to indicate the efficacy of the treatment with an intention-to-treat (ITT) schedule on a carefully chosen personal basis. Despite the diverse cohort, the principal self-similarity allows the self-organized approach [43] which fits the allometric scaling by the fractal structure of the tumor [44]. In consequence, self-organization data mining could prove that the results have the necessary information to measure and evaluate survival [45]. Medical evaluations of survival conventionally apply Kaplan–Meier (KM) non-parametric estimator [46] for incomplete observations. KM is useful to examine the probability of lifetime and effectivity of the chosen treatment for such lethal diseases such as cancer. Taking the self-similarity into consideration, the hazard function must be a self-similar time function [47] and in consequence, the KM could be approached with Weibull distribution [48]
The invariance of magnification (scale invariance, when the up- or down-magnification show similar structures) is the form of self-similarity, which is a typical consequence of the self-organizing processes [49].
No “average” patient exists; the cohort is widely mixed. While in randomized studies, the randomization enables unbiased estimation of treatment effects; observational studies are typically not random. Propensity score matching (PSM) is a method of statistical analysis to estimate the effect of a treatment by accounting for the covariates that predict receiving the treatment. PSM is a conditional probability of being exposed given a set of covariates attempts to reduce the bias by the confounding variables [50]. The PSM improves the evidence level of the observation study, intending to reduce the treatment assignment bias by matching and mimicking randomization, by samples receiving the treatment that is comparable on all observed covariates without receiving the treatment. The possible reference solutions apply proper historical control from the same clinic/hospital where the observational study is performed, retrospectively choosing the same conditions. In the case of mEHT, the PSM was chosen from the patients from the same hospital with the same diseases and stages.
The PSM increases the trustworthiness of the obtained results [50,51] by combining them with an available database, selecting similar cases to be used as a control [52]. Selecting a comparative group of patients uses data mining in large and representative databases, defining the disease’s relevant and characteristic properties and the patients’ conditions.
The expectation that selecting the independent parameters from the actual therapy does not change during the complete curative or palliative process drives the propensity score comparison. The propensity score method gives statistical proof if the confounding variables are chosen well [53]. Advanced cancerous cases may limit the selection because of the large variety of previously failed treatments, so the applied database has to be large enough to mine the appropriate data.
Mathematical/statistical estimates may increase the single-arm’s strength of evidence. The single-arm has complete information about the patients [45,54], but their evaluation is difficult due to the missing comparison cohort. The self-organizing behavior of tumors provides a satisfactory accuracy of evaluating the single-arm, statistically deducting a reference group from the measured data [55].
Repeating the single-arm trial in different research places at different times for the same stage of the disease provides more realistic confidence supporting the evidence. The data pool of the different single-arm studies may increase the evidence level significantly. The most convincing statistical result is the similar, statistically equivalent survival curves of the studies performed at various times in various clinics and countries.

2.2. Clinical Results

The definite primary endpoint of the mEHT studies is the synergy of overall survival (OS) time with quality of life (QoL). Secondary endpoints are the local effects (response, remission, and local control). While the secondary endpoints are popular, they do not provide enough information about the patient’s overall status.

2.2.1. Safety

The Phase 1 clinical trial safety measure was made with patients having advanced glioblastoma multiforme (GBM). This safety study approved the applicability of mEHT to such sensitive organs as the brain without remarkable side effects, even with drastic transcranial dose escalation [56], Table 1. Alkylating chemotherapy (ACNU, nimustin) was administered at a dose of 90   m g / m 2 on day 1 of 42 days for up to 6 cycles or until tumor progression Additional adverse effects of mEHT were not observed.
The GBM is a very fast-growing tumor that spreads rapidly to nearby normal brain tissue, but rarely forms extraneuronal metastases [57], so the distant dissemination to other organs is practically excluded. The selective behavior of the mEHT, and so the strict locality of the heating process, concentrates the energy on the malignant regions and the healthy tissues are unlikely to have harmful doses. Due to the nanoscopic heating, the common problem of the localization of the heat effect [58] is automatically controlled. The adverse effects were measured by dose escalation of mEHT complementary to the ACNU chemotherapy in four groups, increasing the weekly treatment dose from standard two/week to five/week.
The results showed convincing safety of the method, even with extra-large (practically not applied) doses [56]. An essential consequence of this safety trial is when a safe treatment of such a sensitive organ as the brain with such advanced disease as GBM could be performed, we may also expect safety for various other organs.
The radiotherapy combined mEHT trial (n = 20) was safe; no edema appeared with good local control of advanced GBM WHO Grade III–IV.
The optimal dose was determined by dose escalation of mEHT in the Phase I clinical study (n = 19) for patients with relapsed, refractory, or progressive heavily pretreated ovarian cancer. The dose optimum in this disease was 150 W/60 min [59].
Dose escalation of intravenous vitamin C (ivC) together with mEHT was measured with a Phase I safety study, which showed that 1.5 mg/kg ivC is safe for in Stage III–IV non-small cell lung cancer (NSCLC) patients [60].
A metabolically controlled complex therapy package of treatments could be effective in most advanced metastatic cases [61,62,63,64]. The monotherapy application of mEHT also presented promising results for patients with advanced disease when other therapies had failed [65,66].

2.2.2. Survival Time

The clinical results of overall survival (OS) with the synergy of the quality of life (QoL) show the feasibility of complementary applications of mEHT with all standard adjuvant and neoadjuvant oncotherapies, including immuno-oncologic and integrative therapies.
The cervix cancers were studied in a Phase III double-arm randomized prospective controlled trial involving two–two groups in both arms, patients ±HIV [67,68]. Metabolic response (MR) was measured by PET [69]. The advantages of mEHT appeared in the higher complete metabolic response (CMR) of the HIV groups compared to the control groups. The continuation of a Phase III randomized controlled study for advanced cervical cancer patients provides new insights through the strong evidence of mEHT’s efficacy to improve the three-year overall survival [70], Table 2. The OS for patients with FIGO Stage III disease had a significant (p = 0.04) increase with mEHT addition compared to without it. The disease-free survival (DFS) was also significantly longer (p = 0.04) for these patients.
A Phase II randomized double-arm study compared platinum-based chemotherapy to additional complementary mEHT for patients with recurrent cervical cancer, dominantly disseminated squamous cell carcinoma, in a broad range of FIGO statuses [71]. The obtained overall remission rate at seven months of follow-up was significantly better in the active mEHT arm (p = 0.02), and despite of great difference, the overall survival had not reached the significance level (p = 0.24) [72], probably due to the small number (n = 20 + 18) of participants). FIGO Stage 2 locally advanced cervical cancer with lymph node metastases in more than half of the treated patients was studied in retrospective observational studies in a double-arm comparison of standard chemoradiotherapy and its extension with mEHT [73]. The results are convincing (Table 3).
A comparison of the mEHT-treated and nontreated patients with the same stage brain tumors in the same research group in Italy [74] showed a highly significant (p = 0.0006) response rate compared to the conventional control and also for GBM (p = 0.026). The details of the study are shown in Table 4. The theoretical self-similar reference arm corresponds well with the KM plot of the nontreated group of patients [55].
Retrospective observational study with mEHT-treated and untreated arms for advanced pancreatic cancer showed that mEHT was successfully applicable for various pancreatic tumors [75], and also for the specially nonresectable tumors compared to PSM [76], as well as relative long overall survival; Table 5.
A randomized double-arm Phase II study (n = 49 + 48) showed significant improvement in survival and quality of life for patients with Stage III-IV NSCLC treated with mEHT in combination with high-dose vitamin C infusion and providing the best supportive care (BSC) in both arms [79]. The three months follow-up remission rate was significantly better (p = 0.0073) in the active mEHT arm than without it, and simultaneously the survival was also significantly (p < 0.0001) improved. A randomized two arms Phase II clinical trial for advanced non-small-cell lung cancer (NSCLC) patients also showed a clear advantage of mEHT application, showing a significant increase (p < 0.001) in the patients’ OS [79].
The mEHT method also demonstrates the feasibility of treating advanced small-cell lung cancer (SCLC) patients [80], where a significant (p = 0.02) increase in overall survival was observed compared to the control arm of the study.
The mEHT can also be used advantageously in gastrointestinal cases [62,64,81]. The first-line, single-arm, retrospective clinical study (n = 40) of metastatic colon cancer complementary to Bevacizumab+FOLFOX [82] observed progression-free survival (PFS) of 12.1 months and OS 21.4 months, which are remarkably good results.
Neoadjuvant (preoperative) mEHT treatment for locally advanced rectal cancer was studied in a double-arm trial (n = 62 + 58) [83]. The tumor regression was significantly (p = 0.0086) decreased by the tumor volume in the control arm, while in active mEHT treatments, the regression grade was uniform (p = 0.91) and independent of the tumor size. Despite lower radiation doses in the mEHT group, the clinical measures were comparable to the control group; the proportion of downstaging (80.7% vs. 67.2%) and pathologically complete response (pCR, not only for imaging) (17.7% vs. 8.6%) was higher with mEHT than without it. The pathological T-stage (ypT) was significantly (p = 0.049) better with mEHT, and also the rejection margin was significantly (p = 0.013) improved by mEHT application. The survival measures (overall, disease-free, local recurrence-free, metastatic recurrence-free) were all improved by the mEHT, but the differences did not reach a significant level (p = 0.05).
Another Phase II single-arm (n = 60) clinical trial for neoadjuvant mEHT was performed for rectal cancer in cT3-4 or cT2N+ stages [77]. The therapy showed T- and N-downstaging in 40 patients (66.7%) and 53 patients (88.3%), respectively. In total, 15% of patients had complete pathologic response in the T-stage, and 76.7% in the N-stage.
The treatment of peritoneal carcinomatosis with malignant ascites with mEHT combined with traditional Chinese medicine compared to intraperitoneal chemoinfusion (IPCI) in a Phase II randomized double-arm trial [78] observed a better overall response (77.7% in the study arm, while 63.8% in control).
Notably, for cholangiocarcinoma [84,85] and tumors of the hepatopancreatobiliary system [38], treatments with mEHT show the feasibility of the method on presently low-success curing-rate tumor localizations. Two Phase II studies proved the successful applicability of mEHT in hepatocellular carcinoma [86,87].
The mEHT was successfully applied to advanced breast cancer [88], including triple-negative cases [89,90], and leiomyosarcoma of the breast [91]. The study of thirteen patients with complicated, advanced invasive ductal breast carcinoma, mostly triple-negative immunohistochemical status and multiple metastases, showed more than two months of median survival [92].
A study of advanced ovarian cancer treated with mEHT complementary combined with paclitaxel and cisplatin chemotherapy showed less toxicity and adverse effects than cisplatin [93]. The combined two drugs showed no Grade III or IV toxicity and a 57% remission rate during 30 months follow-up, together with 85.7% survival in the same period [94]. A Phase II study of heavily pretreated, mostly platinum-resistant ovarian cancer with relapsed, refractory, or progressive stages treated with mEHT, showed a remarkable 7.5 months median survival in the same treatment period [59].

2.2.3. Comparison of Survival Curves

Glioblastoma treatments with mEHT have numerous single-arm prospective and retrospective clinical studies [74,95,96,97,98,99,100,101,102]. The studies had the same treatment protocol in Figure 2.
Multiple single-arm studies are appropriate for comparison. Evaluating the comparison of these results showed an excellent match with each other, Figure 3. The similarity of the curves may mean strong evidence of the mEHT success.
The various GBM studies following the same general protocol are comparable, Table 6.
A meta-analysis also showed the superiority of the mEHT treatment [104]. Furthermore, the results could be compared with the effective update chemotherapy of GBM, the temozolomide (TMZ) [105], Figure 4. The comparison with the pooled data shows the advantage of mEHT again.
Furthermore, the pool of survival rates of the GBM patients (n = 325) in various single-arm studies showed good agreement with the invasive transcranial brachytherapy ± invasive hyperthermia of the same disease [106], Figure 5. The invasive method (brachytherapy alone or combined with invasive hyperthermia (iHT)) do not differ from the survival results of noninvasive mEHT.
The tumor treating field (TTF) is an emerging electromagnetic therapy for GBM, showing the electric field’s efficacy for this disease [107]. Its focal point is the cytokinetic “neck” at the end of the mitotic spindle, applying nonthermal effects with capacitive coupling [108]. The electric field of TTF reorients the highly polarizable microtubules and actin fibers, and it may arrest the cytoskeleton’s polymerization process and inhibit the assembling of the mitotic spindle [109]. Impressive clinical results were achieved with TTF, proving the feasibility of the nonthermal application of bioelectromagnetic processes against malignant proliferation. A comparative meta-analysis of TTF and mEHT [110] establish that the use of both mEHT and TTF in the treatment of glioblastomas can improve overall survival. A comparison of the mEHT pool and TTF results [111] showed no difference in the clinical results of the two electromagnetic therapies, Figure 6. The differences between TTF and mEHT are basically in the practical application and length to complete the therapy. TTF fixes a hat with multiple electrodes which the patient has to wear 18 h/day for several months, while mEHT makes intensive 60 min treatments every second days in 12 sessions.
The successes of the complex therapy involving mEHT appear as some challenges for randomized studies of GMB [112]. An intensive naturopathic-oncotherapy (Bevacizumab + Boswellia serrata + Curcumin) combination with mEHT looks feasible for GBM patients in a terminal state [113]. This was measured earlier also using high-dose IV Vitamin C (30 g) combined with Thalidomide (50 mg) and Boswellia serrata (400 mg) plus fortecortin (0.5 mg) [114] in cases for patients for whom the chemotherapy is contraindicated.
The GBM has only rarely metastases regardless of the ability of GBM cells to be disseminated over the blood–brain barrier (BBB) [115]. This specialty makes the disease mostly localized, and the various metastases only rarely influence the survival time, which is otherwise the drastic limiting factor of survival in other malignancies. Due to these conditions, we do not expect such unified survival probabilities in other tumors as we observed in GBM.
There are protocols and guidelines not only for the therapy of central nervous system, but those available for various other applied therapies as well [116,117].
We compared the various pancreas studies collected in Table 7. The treatment success of advanced pancreatic cancer has shown relatively little development recently. The applied protocols are that most pancreatic malignancies are non-resectable or can only be partially excised (R1) and frequently make liver metastases, increasing the mortality of this disease. Intensive clinical research with mEHT is underway in this area [61,118,119,120,121]. A comparison of the Kaplan–Meier survival curves shows a definite similarity in the survival probabilities, Figure 7, without such unified curves as we observed in GBM.
Another massively present malignancy in the world is advanced non-small-cell lung cancer (NSCLC). Application of mEHT on advanced NSCLCs also has numerous case studies [63,122,123,124,125,126] and trials [60,79,127,128,129,130]. The comparison of the survival distributions shows similarities again, and the data pool remarkably differs from the propensity score data by mining in USA and EU databases [130], Figure 8.
The comparison of the survival plots from different NSCLC studies Figure 9, Table 8, is not as unified as the GBM comparison in Figure 1, because of the multiple variations of metastases and tumor locations in the lung. However, no significant differences appeared in the observed survival probabilities, so the pooling of data was also possible.
The liver metastases of colorectal carcinoma appear to be a frequent complication in this malignancy. Multiple trials have demonstrated the efficacy of mEHT in this case, too [99,132,133]. The colorectal carcinoma survival studies also may be compared Table 9.
The comparison of the measured Kaplan–Meier non-parametric distributions from different studies, Figure 10, had no significant differences but differed considerably due to the high variability of these metastatic conditions. The pooling here had no statistical basis.

2.2.4. Quality of Life

The synergy of OS with QoL is especially important in the less-successful conventional therapies, such as the brain, pancreas, lung, and liver, which are otherwise frequent metastatic locations from various malignancies when their mortality is exceptionally high.
The Phase I safety study (n = 35, involving Stage IV n = 17) for NSCLC observed adverse effects (fatigue, nausea, vomiting, diarrhea, headache) only rarely and also temporarily [60]. The function subscale of QLQ-C30 scores showed significant improvement in physical status after four weeks of treatments compared to before the therapy and getting gains in all other categories (emotional, cognitive, social, global) without reaching the p < 0.05 significance level. However, the advantage results of mEHT on the symptoms subscale were significant after four weeks in most categories (such as fatigue, dyspnea, insomnia, appetite, and diarrhea). The decrease in nausea/vomiting, pain, and constipation were also observed without significance. The Phase II continuation of the same NSCLC study [79] compared the QoL data in time, not to the baseline. The control arm for comparison was received by randomization of the included cohort of patients. In this randomized study, the QoL significantly improved in all conditional (functional) categories and physiometric (symptoms) categories.
The data in the comparison of IPCI with mEHT favored the latter by improved overall QoL with 32.3% and 49.2%, respectively [78]. Primarily, pain relief in lung cancer was studied in a propensity case-controlled study [124], showing increasing pain after mEHT, which gradually lowered, and after a long time, the effective analgesic score decreased from the original 8.5 to −83.7%, which was more than a 90% improvement compared to baseline before therapy.
Similar results were obtained with less radiation dose combined with mEHT than alone [77,83], which may result in fewer adverse effects and higher QoL of the treated patients.
Improved QoL functions were measured with a Phase III randomized prospective, controlled clinical study of the same cohort of patients showing decreasing toxicity and increased quality of life [134]. Six weeks after the mEHT therapy, the cognitive function significantly improved compared to the control arm. A significant increase in social and emotional function was measured three months post-treatment, while fatigue and pain were significantly decreased simultaneously. Notable the QoL function was also improved and the ratio of mEHT-arm to the control grew by 1.6, 1.3, 2, 2, 10.8, 1.7, and 5.7 for visual analog, global health, pain reduction, nausea/vomiting reduction, fatigue reduction, cognitive functions, emotional functions, role function, and physical status, respectively.

2.2.5. Immunogenic Effects

The immunogenic effects of hyperthermia are currently at the forefront of research and fit well into the general trend that immunology would require. However, hyperthermia applied by itself can also produce immunogenic effects. Cancer patients usually have a weakened immune system that requires stimulation. One of the first series of hyperthermia boosted by immunostimulants was published in 1986 [135]. The double-arm study (n = 77, 75%, surgery, 25% inoperable) of pancreas adenocarcinoma applied capacitive coupling hyperthermia 13.56 MHz plus chemotherapy (doxorubicin 30 mg/m2, 5-FU 500 mg/m2, mitomycin C 5 mg/m2). In total, 42% of patients had chemotherapy in the inoperable group, 5% radiotherapy, 21% radio-chemotherapy, and 32% had no prior therapy due to their refractive status. The immune stimulation used granulocyte–macrophage colony-stimulating factor (GM-CSF) and the results were very encouraging. The ratios of survival percentages between immune-treated and untreated groups showed 2.0-, 5.8-, and 3.0-times increase of 6, 12, and 18 months survival by immune boosting, respectively. The results were highly significant.
The primary task of immunogenic stimulation is to restore homeostatic control and ensure the systemic surveillance of healthy processes. The local treatment becomes systematic in this way and allows attacking the distant micro- and macro-metastases, forming abscopal outcomes [35]. The abscopal principle could be used as a new anti-cancer vaccination strategy with immune stimuli [136], emerging “hyperthermic immunotherapy” [137], and developing tumor-directed immunotherapy [138]. Studies have shown that combining mEHT with traditional Chinese medicine as an immune booster also has abscopal effects [139].
When the patients’ immune system is strong enough to develop tumor-specific immune reactions, then the mEHT works without extra immune stimulation. A Phase III clinical trial for uterus cervix cancer proved the abscopal phenomenon [68,140], forming the complete metabolic response almost five-times more than the otherwise systemic chemotherapy in the control arm. The extra-pelvic response abscopal effect does not depend on the HIV status of the patient. The mEHT treatment of colon cancer clearly shows an abscopal effect in liver metastases [82].
A new therapeutic field is the complementarity application of checkpoint inhibitors (CPIs) combined with mEHT [141,142]. The immune action by CPI leads to a definite abscopal effect in clinical practice [143,144].
Viral immunostimulant with Newcastle viruses is a new, emerging complementary treatment with mEHT, allowing a new strategy for whole-body action [145]. The mEHT has a complex synergy with viral immune stimulation [146,147], also using the ICD process for developing tumor-specific immune activity [148]. The mEHT is a part of multimodal immunotherapy for patients with GBM [149], allowing personalized medicine in glioblastoma multiforme [150]. The complex therapy by dendritic cell vaccines and other immune stimulation to develop ICD within chemotherapy administration might improve the overall survival rate of GBM patients with long-term tumor control [112,151], as well as the induction of ICD during maintenance chemotherapy combined with subsequent multimodal immunotherapy for GBM. A study (n = 41) showed the remarkable benefits of mEHT as part of multimodal immunotherapy for brain tumors in children with DIPG [152], without significant toxicity. The median PFS and OS were 8.4 m and 14.4 m from the time of diagnosis, respectively. The two-year OS was 10.7%. Immunotherapy was applied at the time of progression, when the measured PFS and OS medians were 6.5 m and 9.1 m, respectively.

3. Discussion

The evidence level is improved by comparing the survival data of various studies from different institutes. The relationship between relevant higher evidence-level studies and large databases gives efficacy information. The data of the variant of single-arm observational studies were pooled and used to compare to another type of treatment results. The pooling of data is correct because these survival plots do not vary significantly. The pool of data mimicked the later introduced market surveillance when the various results from very different medical groups were compared. The well-correlating single-arm survival showed that the application of the same protocol at different hospitals, countries, and time provided statistically equivalent data, so it is a usual requirement for worldwide approved drug applications. The technical solution allows easy and safe application [153].
The comparison of Kaplan–Meier survival times of glioblastoma multiforme in six studies made in different institutes gave an excellent agreement between the curves. The match was probably so accurate because the GBM rarely gives non-neural metastases [57], so the distant metastases do not modify the survival. However, the micro- and macro-metastases could critically worsen the survival time. The frequent distant metastases of these tumors are the reason why we do not have perfect equivalence to the KM survival plot in pancreas carcinoma, NSCLC, and colorectal carcinoma, although the differences remain insignificant. The practically identical survival results increase the evidence level of the studies. Due to the insignificant differences, pooled data may be used for comparison with other kinds of cancer treatments for the same tumors. The pooled data showed the superiority of mEHT over temozolomide + radiotherapy treatment for GBM. The mEHT pool agreed well with invasive hyperthermia and the results of other nonthermal bioelectromagnetic therapy (TTF).
The immune stimuli and the immunogenic cancer cell-death [154] are probably a considerable addition to the elongation of the patient’s survival.
The quality-of-life measurements had not revealed extra adverse effects of mEHT and even showed improvements in all functional and physiometric symptomatic scores. The preclinically approved immunogenic effects of mEHT [136,155,156,157], the human response studies [137,141,142], and the observed abscopal effects [138,143,144] probably promote the presented improvement of survival and QoL.
The mEHT, as the synergic therapy of thermal and nonthermal effects, such as other therapies, has limitations. Various challenges appear in general hyperthermia treatments [158,159], which exist and are combined with some specialties in mEHT. The average power of mEHT has to be limited for optimal treatment and the energy absorption has to be balanced, due to:
  • As natural nanoparticles in the membrane, the rafts are heat-sensitive molecular clusters. The too-large absorbed energy destroys the rafts by overheating. The massive distortion of the rafts may degrade the membrane integrity and cause necrosis, losing the apoptotic “harmony” with the homeostasis, which is suboptimal.
  • The selected energy absorption of rafts heats the TME and tissues to a lesser extent. The standard applied SAR in nanoparticles, considering their weight heating, is 0.1–1.5 MW/kg. The approximation of the absorbed power of rafts in selection is SAR > 1 MW/kg [19], which is similar to the standard MNP energies [160]. The increased diffusion redistributes the initial spacing with nanoparticles [161]. The electric field impacts the diffusion of the charged and dipole particles, modifying the electrokinetics of the effusion [162] and the angiogenesis [163]. In case of electric field heating at mEHT, the electrodiffusion modifies the allocation of the gold NP-s too, positions them to the volumes of high electric field, thus promoting the heat on the TME [162]. The heating of the NPs shares the energy, reducing the effect on the membrane rafts, and, despite the increase of temperature, the apoptosis decreases [164]. The distribution of magnetic NP-s, which are modified by the increased diffusion with the temperature [165], has to be impacted too by electrophoresis and electroosmosis, and so the electric field in low frequencies (modulation frequencies of mEHT) regroups them on the same way such as in the case of the non-magnetic metallic NPs [164].
  • The thermal effect happens in nanoscopic local “points”, the rafts. These molecular clusters are sensitive to overheating. When the absorbed energy is too large, it destroys the rafts and the mEHT loses its largest advantage, producing immunogenic cell death (ICD).
  • The large energy absorption extensively forces the spread of heat, and the selection of microscopic differences vanishes. A macroscopic average will characterize the target, and the cellular selection with intended molecular excitations will vanish. The thermal component will become dominant, and the selection mechanisms cannot prevail. A limited thermal component ensures the selection of rafts.
The thermal conditions induce numerous physiological processes interacting with the body’s thermal regulation, which limits the temperature gain (∆T) with the following processes:
  • For human adults the surface heat-loss is floss≅0.15 at rest in the 0 ≤ floss ≤ 1 scale [166], so the heat exchange is intensive enough even by intense local heating. Consequently, the bloodstream in the tumor maintains massive cooling efficacy. The cooling is inhomogeneous; it depends on the vascularization. As such, the thermal factor of mEHT is less homogenous than the nonthermal one.
  • The thermally induced vasoconstriction regulates the blood perfusion and heat conduction in tumors [167,168,169], while the heated healthy tissues in the surroundings have vasodilation. The relative blood flow could promote vascular invasion of the tumor border, reducing the prognostic expectations [170]. The special nano-selectivity and the applied low incident power (about 1/8th of the other conventional local hyperthermia therapies) produce a moderate (fever level) temperature in the tumor mass and its surrounding tissues, which causes much less increase of the blood perfusion than the conventional hyperthermia methods, Figure 11. The thermal damage, which is usually calculated [171], is not considered in mEHT, the nonthermal factor makes the apoptosis, ICD, and immune effects also have a gradient on the tumor border, but it helps focus on the denser tumor by refraction angle.
  • The thermally promoted intensive metabolic activity deprives the ATP sources [172]. However, the massive energy demand of proliferation requests enormous ATP production, which induces anaerobic metabolism, improving the intensive proliferation [173], promoting the malignant processes [174], and leading the growth direction by an acidic invasion front [175].
  • A positive process of ATP deprivation may cause protein aggregation in the cytosol [129,176], destroying the cytoskeleton order. The collapsing cytoskeleton destabilizes the plasma membrane, and the cell necrotizes [176]. Increased temperatures can slow down or even block DNA replication [177], and the DNA strand breaks [178], which completes radiotherapy [179].
  • The intensive thermal effect acts mainly in the S-phase of the cell division cycle [180], while moderate heat shock arrests G1/S and G2/M cell-cycle checkpoints [181]. The nonthermal factor has an intensive block of the last phase of the cell cycle.
Furthermore, the mEHT application must consider the following conditions to avoid the suboptimal treatment:
  • The appropriate frequency is accurately selected around 10 MHz [182,183]. When the frequency is larger (>15 MHz), the membrane impedance becomes too small to select the disordered TME. The current will flow through the entire cell almost equally, neglecting the selection factor of dielectric permittivity.
  • The electromagnetic nano-targeting of rafts has similarities to the molecular targeting of drugs at cancer cells. The chemo dose is limited by poisoning. When the rafts are overheated, the raft protein may coagulate, and no selective heating is possible thereafter.
  • Hyperthermia may reversibly destabilize the raft structures [184], which could mix the time sequences of DAMP production and arrest the immune cells’ activity [185].
  • The approaching of the contact current by mEHT has further limitations. RF safety standards specify the exposure limits [186]. The SAR could be extremely high in the small cross-section when the applicator does not smoothly cover the treatment area and the current flows through a small area which may burn in this touching. The challenge grows when the interface between the skin and electrode has a conductive layer such as sweat, saline, or other aqueous solution. The thin layer may be heated dangerously quickly, so the skin surface must be kept dry.

4. Conclusions

The success of the mEHT therapy of advanced, relapsed, and metastatic tumors has had multiple clinical studies. Some studies have double-arm comparisons, including a Phase III randomized, prospective controlled one for the uterus and cervix. The comparisons show a significant increase in survival time. The Kaplan–Meier survival plots were compared for the single-arm studies. The different studies at various institutes’ and times showed significant correspondence, and the data could be pooled, which increases the evidence level of these observational studies. These clearly show the survival advantage of mEHT. We may conclude that mEHT significantly increases overall survival and the quality of life; consequently, it is a feasible treatment for the presented malignant tumors.

Author Contributions

Conceptualization, S.-Y.L.; methodology, S.-Y.L.; resources, S.-Y.L. and A.M.S.; writing—original draft preparation, S.-Y.L.; writing—review and editing, G.L., L.G. and A.M.S.; supervision, L.G. and A.M.S. All authors have read and agreed to the published version of the manuscript.

Funding

NVKP_16-1-2016-0042 and KDP-2020, both supported by the Ministry of Innovation and Technology, and the National Research, Development and Innovation Fund of Hungary.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Duffell, E. Curative power of fever. Lancet 2001, 358, 1276. [Google Scholar] [CrossRef] [PubMed]
  2. Szasz, O.; Szasz, A. Approaching complexity: Hyperthermia dose and its possible measurement in oncology. Op. J. Biophys. 2021, 11, 68–132. [Google Scholar] [CrossRef]
  3. Seel, M.; Ladik, J. The Tragicomedy of Modern Theoretical Biology, Chapter 1. In Advances in Quantum Chemistry; Elsevier: Amsterdam, The Netherlands, 2019; Volume 81, pp. 1–13. [Google Scholar] [CrossRef]
  4. Vaupel, P.; Hammersen, F. Mikrozirkulation in Malignen Tumoren. In 6. Jahrestagung der Gesellschaft für Mikrozirkulation E.V., München; Karger: Basel, Switzerland, 1982. [Google Scholar]
  5. Charkoudian, N. Skin Blood Flow in Adult Human Thermoregulation: How It Works, When It Does Not, and Why. Mayo Clin. Proc. 2003, 78, 603–612. [Google Scholar] [CrossRef] [PubMed]
  6. Szasz, A. Stimulation and control of homeostasis. Open J. Biophys. 2022, 12, 89–131. [Google Scholar] [CrossRef]
  7. Lee, S.-Y.; Fiorentini, G.; Szasz, A.M.; Szigeti, G.; Szasz, A.; Minnaar, C.A. Quo vadis oncological hyperthermia (2020)? Front. Oncol. 2020, 10, 1690. [Google Scholar] [CrossRef]
  8. Szasz, A. Thermal and nonthermal effects of radiofrequency on living state and applications as an adjuvant with radiation therapy. J. Radiat. Cancer Res. 2019, 10, 1–17. [Google Scholar] [CrossRef]
  9. Szasz, A. Heterogeneous heat absorption is complementary to radiotherapy. Cancers 2022, 14, 901. [Google Scholar] [CrossRef]
  10. Minnaar, C.A.; Szasz, A. Forcing the antitumor effects of HSPs using a modulated electric field. Cells 2022, 11, 1838. [Google Scholar] [CrossRef]
  11. Minnaar, C.A.; Szasz, A.; Lee, S.Y.; Szigeti, G.P.; Szasz, A.M.; Mathe, D. Supportive and palliative care in cancer therapies—Path from tumor-driven therapies to patient-driven ones. Int. J. Clin. Med. 2022, 13, 287–359. [Google Scholar] [CrossRef]
  12. Lee, S.Y.; Szasz, A. Immunogenic Effect of Modulated Electro-Hyperthermia (mEHT) in Solid Tumors. In Interdisciplinary Cancer Research; Springer: Cham, Switzerland, 2022. [Google Scholar] [CrossRef]
  13. Muftuler, T.L.; Hamamura, M.J.; Birgul, O.; Nalcioglu, O. In Vivo MRI Electrical Impedance Tomography (MREIT) of Tumors. Technol. Cancer Res. Treat. 2006, 5, 381–387. [Google Scholar]
  14. Smith, S.R.; Foster, K.R.; Wolf, G.L. Dielectric Properties of VX-2 Carcinoma Versus Normal Liver Tissue. IEEE Trans. Biomed. Eng. 1986, 33, 522–524. [Google Scholar] [CrossRef] [PubMed]
  15. Joy, M.; Scott, G.; Henkelman, M. In vivo detection of applied electric currents by magnetic resonance imaging. Magn. Reson. Imaging 1989, 7, 49–54. [Google Scholar] [CrossRef] [PubMed]
  16. Loewenstein, W.R.; Kanno, Y. Intercellular communication and tissue growth. J. Cell Biol. 1967, 33, 225–234. [Google Scholar] [CrossRef]
  17. Foster, K.R.; Schepps, J.L. Dielectric properties of tumor and normal tissues at radio through microwave frequencies. J. Microw. Power 1981, 16, 107–119. [Google Scholar] [CrossRef] [PubMed]
  18. Szentgyorgyi, A. Electronic Biology and Cancer; Marcel Dekkerm: New York, NY, USA, 1998. [Google Scholar]
  19. Papp, E.; Vancsik, T.; Kiss, E.; Szasz, O. Energy absorption by the membrane rafts in the modulated electro-hyperthermia (mEHT). Open J. Biophys. 2017, 7, 216–229. [Google Scholar] [CrossRef]
  20. Baish, J.W.; Jain, R.K. Fractals and Cancers. Cancer Res. 2000, 60, 3683–3688. [Google Scholar]
  21. Andocs, G.; Rehman, M.U.; Zhao, Q.L.; Papp, E.; Kondo, T.; Szasz, A. Nanoheating without Artificial Nanoparticles Part II. Experimental support of the nanoheating concept of the modulated electro-hyperthermia method, using U937 cell suspension model. Biol. Med. 2015, 7, 247. [Google Scholar] [CrossRef]
  22. Andocs, G.; Rehman, M.U.; Zhao, Q.-L.; Tabuchi, Y.; Kanamori, M.; Kondo, T. Comparison of biological effects of modulated electro-hyperthermia and conventional heat treatment in human lymphoma U937 cell. Cell Death Discov. 2016, 2, 16039. [Google Scholar] [CrossRef]
  23. Martinsen, O.G.; Grimnes, S.; Schwan, H.P. Interface Phenomena and Dielectric Properties of Biological Tissue. In Encyclopedia of Surface and Colloid Science; CRC: Boca Raton, Fl, USA, 2002; pp. 2643–2652. [Google Scholar]
  24. Schwan, H.P.; Takashima, S. Dielectric behavior of biological cells and membranes. Bull. Inst. Chem. Res. Kyoto Univ. 1991, 69, 459–475. [Google Scholar]
  25. Goldberger, A.L.; Amaral, L.A.N.; Hausdorff, J.M.; Ivanov, P.C.; Peng, C.-K. Fractal dynamics in physiology: Altera-tions with disease and aging. Proc. Natl. Acad. Sci. USA 2001, 99 (Suppl. S1), 2466–2472. [Google Scholar] [CrossRef]
  26. Garg, A.D.; Nowis, D.; Golab, J.; Vandenabeele, P.; Krysko, D.V.; Agostinis, P. Immunogen-ic cell death, DAMPs and anticancer therapeutics: An emerging amalgamation. Biochim. Biophys. Acta 2010, 1805, 53–71. [Google Scholar] [PubMed]
  27. Galluzzi, L.; Buqué, A.; Keep, O.; Zitvogel, L.; Kroemer, G. Immunogenic cell death in cancer and infectious disease. Nat. Rev. Immunol. 2017, 17, 97–111. [Google Scholar] [CrossRef] [PubMed]
  28. Obeid, M.; Tesniere, A.; Ghiringhelli, F.; Fimia, G.M.; Apetoh, L.; Perfettini, J.-L.; Castedo, M.; Mignot, G.; Panaretakis, T.; Casares, N.; et al. Calreticulin exposure dictates the immunogenicity of cancer cell death. Nat. Med. 2007, 13, 54–61. [Google Scholar] [CrossRef]
  29. Multhoff, G.; Botzler, C.; Wiesnet, M.; Müller, E.; Meier, T.; Wilmanns, W.; Issels, R.D. A stress-inducible 72-kDa heat-shock protein (HSP72) is expressed on the surface of human tumor cells, but not on normal cells. Int. J. Cancer 1995, 61, 272–279. [Google Scholar] [CrossRef] [PubMed]
  30. Multhoff, G.; Botzler, C.; Jennen, L.; Schmidt, J.; Ellwart, J.; Issels, R. Heat shock protein 72 on tumor cells: A recognition structure for natural killer cells. J. Immunol. 1997, 158, 4341–4350. [Google Scholar] [CrossRef] [PubMed]
  31. Vancsik, T.; Mathe, D.; Horvath, I.; Várallyaly, A.A.; Benedek, A.; Bergmann, R.; Krenács, T.; Benyó, Z.; Balogh, A. Modulated electro-hyperthermia facilitates NK-cell infiltration and growth arrest of human A2058 melanoma in a xenograft model. Front. Oncol. 2021, 11, 590764. [Google Scholar] [CrossRef]
  32. Klune, J.R.; Dhuper, R.; Cardinal, J.; Billiar, T.R.; Tsung, A. HMGB1: Endogenous Danger Signaling. Mol. Med. 2008, 14, 476–484. [Google Scholar] [CrossRef]
  33. Asadzadeh, Z.; Safarzadeh, E.; Safaei, S.; Baradaran, A.; Mohammadi, A.; Hajiasgharzadeh, K.; Derakhshani, A.; Argentiero, A.; Silvestris, N.; Baradaran, B. Current Approaches for Combination Therapy of Cancer: The Role of Immunogenic Cell Death. Cancers 2020, 12, 1047. [Google Scholar] [CrossRef]
  34. Medina, C.B.; Ravichandran, K.S. Do not let death do us part: ‘find-me’ signals in communication between dying cells and the phagocytes. Cell Death Differ. 2016, 23, 979–989. [Google Scholar] [CrossRef]
  35. Szasz, O. Local Treatment with Systemic Effect: Abscopal Outcome. In Challenges and Solutions of Oncological Hyperthermia; Szasz, A., Ed.; Cambridge Scholars: Newcastle upon Tyne, UK, 2020; Volume 11, pp. 192–205. Available online: https://www.cambridgescholars.com/challenges-and-solutions-of-oncological-hyperthermia (accessed on 9 October 2020).
  36. Krenacs, T.; Meggyeshazi, N.; Forika, G.; Kiss, E.; Hamar, P.; Szekely, T.; Vancsik, T. Modulated electro-hyperthermia-induced tumor damage mechanisms revealed in cancer models. Int. J. Mol. Sci. 2020, 21, 6270. [Google Scholar] [CrossRef]
  37. Szasz, A.M.; Minnaar, C.A.; Szentmartoni, G.; Szigeti, G.P.; Dank, M. Review of the clinical evidences of modulated electro-hyperthermia (mEHT) method: An update for the practicing oncologist. Front. Oncol. 2019, 9, 1012. [Google Scholar] [CrossRef]
  38. Herold, Z.; Szasz, A.M.; Dank, M. Evidence based tools to improve efficiency of currently administered oncotherapies for tumors of the hepatopancreatobiliary system. World J. Gastrointest. Oncol. 2021, 15, 1109–1120. [Google Scholar] [CrossRef] [PubMed]
  39. Lee, S.Y.; Kim, J.-H.; Han, Y.-H.; Cho, D.H. The effect of modulated electro-hyperthermia on temperature and blood flow in human cervical carcinoma. Int. J. Hyperth. 2018, 34, 953–960. [Google Scholar] [CrossRef] [PubMed]
  40. Lee, S.Y.; Kim, M.-G. The effect of modulated electro-hyperthermia on the pharmacokinetic properties of nefopam in healthy volunteers: A andomized, single-dose, crossover open-label study. Int. J. Hyperth. 2015, 31, 869–874. [Google Scholar] [CrossRef] [PubMed]
  41. Lee, S.Y.; Kim, M.-G. Effect of modulated electrohyperthermia on the pharmacokinetics of oral transmucosal fentanyl citrate in healthy volunteers. Clin. Ther. 2016, 38, 2548–2554. [Google Scholar] [CrossRef]
  42. Institute of Medicine (US) Committee on Technological Innovation in Medicine; Gelijns, A. Comparing the Development of Drugs, Devices, and Clinical Procedures; National Academies Press (US): Washington, DC, USA, 1990. [Google Scholar]
  43. Bru, A.; Albertos, S.; Subiza, J.L.; García-Asenjo, J.L.; Bru, I. The universal dynamics of tumor growth. Biophys. J. 2003, 85, 2948–2961. [Google Scholar] [CrossRef]
  44. Guiot, C.; Degiorgis, P.G.; Delsanto, P.P.; Gabriele, P.; Deisboeck, T.S. Does tumor growth follow a ‘universal law’? J. Theor. Biol. 2003, 225, 147–151. [Google Scholar] [CrossRef]
  45. Szasz, O.; Szasz, A.M.; Szigeti, G.P.; Szasz, A. Data mining and evaluation of single arm clinical studies, Chapter 2. In Recent Developments in Engineering Research; Yong, X., Ed.; GAN Publishing: London, UK, 2020; Volume 3, pp. 15–74. [Google Scholar]
  46. Etikan, I.; Abubakar, S.; Alkassim, R. The Kaplan Meier Estimate in Survival Analysis. Biom. Biostat. Int. J. 2017, 5, 00128. [Google Scholar] [CrossRef]
  47. Szasz, O.; Szigeti, G.P.; Szasz, A. On the Self-Similarity in Biological Processes. Open J. Biophys. 2017, 7, 183–196. [Google Scholar] [CrossRef]
  48. González, M.M.; Joa, J.A.G.; Cabrales, L.E.B.; Pupo, A.E.B.; Schneider, B.; Kondakci, S.; Ciria, H.M.C. Is cancer a pure growth curve or does it follow a kinetics of dynamical structural transformation? BMC Cancer 2017, 17, 174. [Google Scholar] [CrossRef]
  49. Kurakin, A. The self-organizing fractal theory as a universal discovery method: The phenomenon of life. Theor. Biol. Med. Model. 2011, 8, 4. [Google Scholar] [CrossRef] [PubMed]
  50. Rosenbaum, P.R.; Rubin, D.B. The Central Role of the Propensity Score in Observational Studies for Causal Effects. Biometrika 1983, 70, 41–55. [Google Scholar] [CrossRef]
  51. Rosenbaum, P.R.; Rubin, D.B. Reducing Bias in Observational Studies Using Subclassification on the Propensity Score. J. Am. Stat. Assoc. 1984, 79, 516–524. [Google Scholar] [CrossRef]
  52. Rosenbaum, P.R.; Rubin, D.B. Constructing a Control Group Using Multivariate Matched Sampling Methods That Incorporate the Propensity Score. Am. Stat. 1985, 39, 33–38. [Google Scholar]
  53. Cochran, W.G. The Effectiveness of Adjustment by Sub-Classification in Removing Bias in Observational Studies. Biometrics 1968, 24, 295–313. [Google Scholar] [CrossRef]
  54. Szasz, O.; Szasz, A. Parametrization of survival measures, Part I: Consequences of self-organizing. Int. J. Clin. Med. 2020, 11, 316–347. [Google Scholar] [CrossRef]
  55. Szasz, A.; Szigeti, G.P.; Szasz, A.M. Parametrization of survival measures, Part III: Clinical evidences in single arm studies with endpoint of overall survival. Int. J. Clin. Med. 2020, 11, 389–419. [Google Scholar] [CrossRef]
  56. Wismeth, C.; Dudel, C.; Pascher, C.; Ramm, P.; Pietsch, T.; Hirschmann, B.; Reinert, C.; Proescholdt, M.; Rümmele, P.; Schuierer, G.; et al. Transcranial electro-hyperthermia combined with alkylating chemotherapy in patients with relapsed high-grade gliomas—Phase I clinical results. J. Neurooncol. 2010, 98, 395–405. [Google Scholar] [CrossRef]
  57. Seo, Y.J.; Cho, W.H.; Kang, D.W.; Cha, S.H. Extraneural metastasis of glioblastoma multiforme presenting as an unusual neck mass. J. Korean Neurosurg. Soc. 2012, 51, 147–150. [Google Scholar] [CrossRef]
  58. Singh, M.; Singh, T.; Soni, S. Pre-operative Assessment of Ablation Margins for Variable Blood Perfusion Metrics in a Magnetic Resonance Imaging Based Complex Breast Tumour Anatomy: Simulation Paradigms in Thermal Therapies. Comput. Methods Programs Biomed. 2021, 198, 105781. [Google Scholar] [CrossRef]
  59. Yoo, H.J.; Lim, M.C.; Seo, S.-S.; Kang, S.; Joo, J.; Park, S.Y. Phase I/II clinical trial of modulated electro-hyperthermia treatment in patients with relapsed, refractory or progressive heavily treated ovarian cancer. Jpn. J. Clincal Oncol. 2019, 49, 832–838. [Google Scholar] [CrossRef] [PubMed]
  60. Ou, J.; Zhu, X.; Lu, Y.; Zhao, C.; Zhang, H.; Wang, X.; Gui, X.; Wang, J.; Zhang, X.; Zhang, T.; et al. The safety and pharmacokinetics of high dose intravenous ascorbic acid synergy with modulated electrohyperthermia in Chinese patients with stage III-IV non-small cell lung cancer. Eur. J Pharm. Sci. 2017, 109, 412–418. [Google Scholar] [CrossRef] [PubMed]
  61. Iyikesici, M.S. Long-term survival outcomes of metabolically supported chemotherapy with Gemcitabine-based or FOLFIRINOX regimen combined with Ketogenic diet, hyperthermia, and hyperbaric oxygen therapy in metastatic pancreatic cancer. Complement. Med. Res. 2020, 27, 31–39. [Google Scholar] [CrossRef] [PubMed]
  62. Iyikesici, M.S. Survival outcomes of metabolically supported chemotherapy combined with Ketogenic diet, hyperthermia, and hyperbaric oxygen therapy in advanced gastric cancer. Niger. J. Clin. Pract. 2020, 23, 734–740. [Google Scholar] [CrossRef]
  63. Iyikesici, M.S. Feasibility study of metabolically supported chemotherapy with weekly carboplatin/paclitaxel combined with ketogenic diet, hyperthermia and hyperbaric oxygen therapy in metastatic non-small cell lung cancer. Int. J. Hyperth. 2019, 36, 445–454. [Google Scholar] [CrossRef]
  64. Iyikesici, M.S.; Slocum, A.; Turkmen, E.; Akdemir, O.; Slocum, A.K.; Berkarda, F.B. Complete response of locally advanced (stage III) rectal cancer to metabolically supported chemoradiotherapy with hyperthermia. Int. J. Cancer Res. Mech. 2016, 2, 1–4. [Google Scholar]
  65. Jeung, T.S.; Ma, S.Y.; Yu, J.; Lim, S. Cases that respond to oncothermia monotherapy. Conf. Pap. Med. 2013, 2013, 392480. [Google Scholar] [CrossRef]
  66. Minnaar, C.A.; Szigeti, G.P.; Szasz, A.M.; Kotzen, J.A. Review on the use of modulated electro-hyperthermia as a stand-alone therapy in a palliative setting: Potential for further research? J. Cancer Ther. 2022, 13, 362–377. [Google Scholar] [CrossRef]
  67. Minnaar, C.; Baeyens, A.; Kotzen, J.; Vangu, M.D. Survival of cervical cancer patients with or without associated HIV infection and treated with modulated electro-hyperthermia combined with chemo-radiotherapy, 32nd Annual Meeting of the European Hyperthermia Society, OP 13. Strahlenther. Onkol. 2018, 194, 476. [Google Scholar]
  68. Minnaar, C.A.; Kotzen, J.A.; Ayeni, O.A.; Naidoo, T.; Tunmer, M.; Sharma, V.; Baeyens, A. The effect of modulated electro-hyperthermia on local disease control in HIV-positive and -negative cervical cancer women in South Africa: Early results from a phase III randomized controlled trial. PLoS ONE 2019, 14, e0217894. [Google Scholar] [CrossRef]
  69. Minnaar, C.A.; Baeyens, A.; Ayeni, O.A.; Kotzen, J.A.; Vangu, M.-D. Defining characteristics of nodal disease on PET/CT scans in patients with HIV-positive and -negative locally advanced cervical cancer in South Africa. Tomography 2019, 5, 339–345. [Google Scholar] [CrossRef] [PubMed]
  70. Minnaar, C.A.; Maposa, I.; Kotzen, J.A.; Baeyens, A. Effects of modulated electro-hyperthermia (mEHT) on two and three year survival of locally advanced cervical cancer patients. Cancers 2022, 14, 656. [Google Scholar] [CrossRef] [PubMed]
  71. Lee, S.-Y.; Lee, N.-R.; Cho, D.-H.; Kim, J.S. Treatment outcome analysis of chemotherapy combined with modulated electro-hyperthermia compared with chemotherapy alone for recurrent cervical cancer, following irradiation. Oncol. Lett. 2017, 14, 73–78. [Google Scholar] [CrossRef] [PubMed]
  72. Lee, S.-Y. Concurrent Chemo-Hyperthermia for Recurrent Cervical Cancer after Previous CCRT. In Challenges and Solutions of Oncological Hyperthermia; Szasz, A., Ed.; Cambridge Scholars: Newcastle upon Tyne, UK, 2020; Volume 9, pp. 163–186. [Google Scholar]
  73. Lee, S.Y.; Lee, D.H.; Cho, D.-H. Modulated electrohyperthermia in locally advanced cervical cancer: Results of an observa-tional study of 95 patients. Medicine 2023, 102, e32727. [Google Scholar] [CrossRef] [PubMed]
  74. Fiorentini, G.; Sarti, D.; Casadei, V.; Milandri, C.; Dentico, P.; Mambrini, A.; Guadagni, S. Modulated Electro-Hyperthermia for the Treatment of Relapsed Brain Gliomas. In Challenges and Solutions of Oncological Hyperthermia; Szasz, A., Ed.; Cambridge Scholars: Newcastle upon Tyne, UK, 2020; Volume 6, pp. 110–125. [Google Scholar]
  75. Fiorentini, G.; Sarti, D.; Casadei, V.; Milandri, C.; Dentico, P.; Mambrini, A.; Nani, R.; Fiorentini, C.; Guadagni, S. Modulated electro-hyperthermia as palliative treatment for pancreas cancer: A retrospective observational study on 106 patients. Integr. Cancer Ther. 2019, 18, 1534735419878505. [Google Scholar] [CrossRef]
  76. Petenyi, F.G.; Garay, T.; Muhl, D.; Izso, B.; Karaszi, A.; Borbenyi, E.; Herold, M.; Herold, Z.; Szasz, A.M.; Dank, M. Modulated electro-hyperthermic (mEHT) treatment in the therapy of inoperable pancreatic cancer patients—A single-center case-control study. Diseases 2021, 9, 81. [Google Scholar] [CrossRef]
  77. You, S.H.; Kim, S. Feasibility of modulated electro-hyperthermia in preoperative treatment for locally-advanced rectal cancer: Early phase 2 clinical results. Neoplasma 2020, 67, 677–683. [Google Scholar] [CrossRef]
  78. Pang, C.L.; Zhang, X.; Wang, Z.; Ou, J.; Lu, Y.; Chen, P.; Zhao, C.; Wang, X.; Zhang, H.; Roussakow, S.V. Local modulated electro-hyperthermia in combination with traditional Chinese medicine vs. intraperitoneal chemoinfusion for the treatment of peritoneal carcinomatosis with malignant ascites: A phase II randomized trial. Mol. Clin. Oncol. 2017, 6, 723–732. [Google Scholar] [CrossRef]
  79. Ou, J.; Zhu, X.; Chen, P.; Du, Y.; Lu, Y.; Peng, X.; Bao, S.; Wang, J.; Zhang, X.; Zhang, T.; et al. A randomized phase II trial of best supportive care with or without hyperthermia and vitamin C for heavily pretreated, advanced, refractory non-small-cell lung cancer. J. Adv. Res. 2020, 24, 175–182. [Google Scholar] [CrossRef]
  80. Lee, D.Y.; Haam, S.J.; Kim, T.H.; Lim, J.Y.; Kim, E.J.; Kim, N.Y. Oncothermia with chemotherapy in the patients with Small Cell Lung Cancer. Hindawi Publishing Corporation. Conf. Pap. Med. 2013, 2013, 910363. [Google Scholar]
  81. Garay, T.; Kiss, E.; Szentmartoni, G.; Borbényi, E.; Mühl, D.; Karászi, Á.; Désfalvi, J.; Mohácsi, R.; Kvasnika, M.; Szasz, A.M.; et al. Gastrointestinal Cancer Series Treated with Modulated Electro-Hyperthermia (mEHT)—A Single Centre Experience. In Challenges and Solutions of Oncological Hyperthermia; Szasz, A., Ed.; Cambridge Scholars: Newcastle upon Tyne, UK, 2020; Volume 8, pp. 159–162. [Google Scholar]
  82. Ranieri, G.; Laface, C.; Laforgia, M.; De Summa, S.; Porcelli, M.; Macina, F.; Ammendola, M.; Molinari, P.; Lauletta, G.; Di Palo, A.; et al. Bevacizumab plus FOLFOX-4 combined with deep electro-hyperthermia as first-line therapy in metastatic colon cancer: A pilot study. Front. Oncol. 2020, 10, 590707. [Google Scholar] [CrossRef] [PubMed]
  83. Kim, S.; Lee, J.H.; Cha, J.; You, S.H. Beneficial effects of modulated electro-hyperthermia during neoadjuvant treatment for locally advanced rectal cancer. Int. J. Hyperth. 2021, 38, 144–151. [Google Scholar] [CrossRef] [PubMed]
  84. Marangos, M.; Danilidis, L.; Vakalis, I.; Kouridakis, P.; Papastavrou, A.; Barich, A. Inoperable multifocal intrahepatic cholangiocarcinoma treated with Hyperthermia, IV Vitamin C and ozonated blood autotransfusion Oncothermia. Oncothermia J. 2017, 20, 236–248. [Google Scholar]
  85. Reimnitz, U. Cholangiocellular carcinomas: Survival without symptoms with hyperthermia—A case study. Oncothermia J. 2010, 1, 20–22. [Google Scholar]
  86. Gadaleta-Caldarola, G.; Infusino, S.; Galise, I.; Ranieri, G.; Vinciarelli, G.; Fazio, V.; Divella, R.; Daniele, A.; Filippelli, G.; Gadaleta, C.D. Sorafenib and locoregional deep electro-hyperthermia in advanced hepatocellular carcinoma. A phase II study. Oncol. Lett. 2014, 8, 1783–1787. [Google Scholar] [CrossRef] [PubMed]
  87. Ferrari, V.D.; De Ponti, S.; Valcamonico, F.; Amoroso, V.; Grisanti, S.; Rangoni, G.; Marpicati, P.; Vassalli, L.; Simoncini, E.; Marini, G. Deep electro-hyperthermia (EHY) with or without thermo-active agents in patients with advanced hepatic cell carcinoma: Phase II study. J. Clin. Oncol. 2007, 25, 15168. [Google Scholar] [CrossRef]
  88. Nagata, T.; Kanamori, M.; Sekine, S.; Arai, M.; Moriyama, M.; Fujii, T. Clinical study of modulated electro-hyperthermia for advanced metastatic breast cancer. Mol. Clin. Oncol. 2021, 14, 103. [Google Scholar] [CrossRef] [PubMed]
  89. Garay, T.; Borbényi, E.; Szasz, A.M.; Kulka, J.; Madaras, L.; Somorácz, A.; Lóránt, G.; Molnar, B.A.; Gyorke, T.; Galgoczy, H.; et al. Challenges and Solutions of Oncological Hyperthermia; Szasz, A., Ed.; Cambridge Scholars: Newcastle upon Tyne, UK, 2020; Volume 10, pp. 187–191. [Google Scholar]
  90. Iyikesici, M.S.; Slocum, A.K.; Slocum, A.; Berkarda, F.B.; Kalamian, M.; Seyfried, T.N. Efficacy of metabolically supported chemotherapy combined with ketogenic diet, hyperthermia, and hyperbaric oxygen therapy for stage IV triple-negative breast cancer. Cureus 2017, 9, e1445. [Google Scholar] [CrossRef]
  91. Lee, S.Y.; Lee, N.-R. Positive response of a primary leiomyosarcoma of the breast following salvage hyperthermia and pazopanib. Korean J. Intern. Med. 2018, 33, 442–445. [Google Scholar] [CrossRef]
  92. Szasz, A.M.; Szentmartoni, G.; Garay, T.; Borbényi, E.; Mohacsi, R.; Kulka, J.; Madaras, L.; Kovács, K.A.; Lóránt, G.; Molnár, B.Á.; et al. Breast Cancer Series Treated with Modulated Electro-Hyperthermia (mEHT)—A Single Centre Experience. In Challenges and Solutions of Oncological Hyperthermia; Szasz, A., Ed.; Cambridge Scholars: Newcastle upon Tyne, UK, 2020; Volume 5, pp. 105–109. [Google Scholar]
  93. Kim, K.; Kim, J.-H.; Kim, S.C.; Kim, Y.B.; Nam, B.-H.; No, J.H.; Cho, H.; Ju, W.; Suh, D.H.; Kim, Y.H. Modulated electro-hyperthermia with weekly paclitaxel or cisplatin in patients with recurrent or persistent epithelial ovarian, fallopian tube or primary peritoneal carcinoma: The KGOG 3030 trial. Exp. Ther. Med. 2021, 22, 787. [Google Scholar] [CrossRef]
  94. Deniz, G.I.; Can, A.; Tansan, S. Chemotherapy and radiofrequency hyperthermia in advanced ovarian cancer. J. Clin. Oncol. 2022, 40 (Suppl. S16), e17550. [Google Scholar] [CrossRef]
  95. Sahinbas, H.; Groenemeyer, D.H.W.; Boecher, E.; Szasz, A. Retrospective clinical study of adjuvant electro-hyperthermia treatment for advanced brain-gliomas. Dtsch. Z. Fuer Onkol. 2007, 39, 154–160. [Google Scholar] [CrossRef]
  96. Fiorentini, G.; Sarti, D.; Milandri, C.; Dentico, P.; Mambrini, A.; Fiorentini, C.; Mattioli, G.; Casadei, V.; Guadagni, S. Modulated electrohyperthermia in integrative cancer treatment for relapsed malignant glioblastoma and astrocytoma: Retrospective multicenter controlled study. Integr. Cancer Ther. 2018, 18, 1534735418812691. [Google Scholar] [CrossRef] [PubMed]
  97. Douwes, F.; Douwes, O.; Migeod, F.; Grote, C.; Bogovic, J. Hyperthermia in Combination with ACNU Chemotherapy in the Treatment of Recurrent Glioblastoma; St. Georg Klinik: Hamburg, Germany, 2006. [Google Scholar]
  98. Fiorentini, G.; Sarti, D.; Milandri, C.; Dentico, P.; Mambrini, A.; Guadagni, S. Retrospective observational clinical study on relapsed malignant gliomas treated with electro-hyperthermia. Int. J. Neurooncol. Brain Tumors 2017, 1, 9–13. [Google Scholar]
  99. Parmar, G.; Rurak, E.; Elderfield, M.; Li, K.; Soles, S.; Rinas, A. 8-Year Observational Study on Naturopathic Treatment with Modulated Electro-Hyperthermia (mEHT): A Single-Centre Experience. In Challenges and Solutions of Oncological Hyperthermia; Szasz, A., Ed.; Cambridge Scholars: Newcastle upon Tyne, UK, 2020; Volume 13, pp. 227–266. [Google Scholar]
  100. Solodkiy, V.A.; Panshin, G.A.; Izmailov, T.R.; Shevchenko, T.A. The first experience of application of remote radiotherapy in combination with hyperthermia (oncothermia) in the treatment of patients with primary gliomas of the brain of a high degree of malignancy. Boпрoсы Онкoлoгии (Probl. Oncol.) 2021, 67, 2. (In Russian) [Google Scholar] [CrossRef]
  101. Hager, E.D.; Sahinbas, H.; Groenemeyer, D.H.; Migeod, F. Prospective phase II trial for recurrent high-grade malignant gliomas with capacitive coupled low radiofrequency (LRF) deep hyperthermia. ASCO. J. Clin. Oncol. 2008, 26, 2047. [Google Scholar] [CrossRef]
  102. Hager, E.D.; Dziambor, H.; App, E.M.; Popa, C.; Popa, O.; Hertlein, M. The treatment of patients with high-grade malignant gliomas with RF-hyperthermia. Proc. Am. Soc. Clin. Oncol. 2003, 22, 1–5. [Google Scholar]
  103. Szasz, A.; Dani, A.; Varkonyi, A.; Magyar, T. Retrospective Analysis of 1180 Oncological Patients Treated by Electro-Hyperthermia in Hungary. In Proceedings of the Jahreskongress der Deutschen Gesellschaft für Radioonkologie, Karlsruhe, Germany, 26–29 May 2005. DEGRO 11. [Google Scholar]
  104. Roussakow, S. Clinical and economic evaluation of modulated electrohyperthermia concurrent to dose-dense temozolomide 21/28 days regimen in the treatment of recurrent glioblastoma: A retrospective analysis of a two-centre German cohort trial with systematic comparison and effect-to-treatment analysis. BMJ Open 2017, 7, e017387. [Google Scholar] [CrossRef]
  105. Stupp, R.; Mason, W.P.; van den Bent, M.J.; Weller, M.; Fisher, B.; Taphoorn, M.J.; Belanger, K.; Brandes, A.A.; Marosi, C.; Bogdahn, U.; et al. European Organisation for Research and Treatment of Cancer Brain Tumor and Radiotherapy Groups; National Cancer Institute of Canada Clinical Trials Group; Radiotherapy plus concomitant and adjuvant temozolomide for glioblastoma. N. Engl. J. Med. 2005, 352, 987–996. [Google Scholar] [CrossRef]
  106. Sneed, P.K.; Stauffer, P.R.; McDermott, M.W.; Lamborn, K.R.; Weaver, K.A.; Prados, M.D.; Chang, S.; Malec, M.K.; Spry, L.; Malec, M.K.; et al. Survival benefit of hyperthermia in a prospective randomized trial of brachytherapy boost +/− hyperthermia for glioblastoma multiforme. Int. J. Radiat. Oncol. Biol. Phys. 1998, 40, 287–295. [Google Scholar] [CrossRef]
  107. Wong, E.T.; Lok, E.; Swanson, K.D. An evidence-based review of alternating electric fields therapy for malignant gliomas. Curr. Treat. Options Oncol. 2015, 16, 40. [Google Scholar] [CrossRef] [PubMed]
  108. Kirson, E.D.; Gurvich, Z.; Schneiderman, R.; Dekel, E.; Itzhaki, A.; Wasserman, Y.; Schatzberger, R.; Palti, Y. Disruption of Cancer Cell Replication by Alternating Electric Fields. Cancer Res. 2004, 64, 3288–3295. [Google Scholar] [CrossRef]
  109. Kirson, E.D.; Dbaly, V.; Tovarys, F.; Vymazal, J.; Soustiel, J.F.; Itzhaki, A.; Mordechovich, D.; Steinberg-Shapira, S.; Gurvich, Z.; Schneiderman, R.; et al. Alternating electric fields arrest cell proliferation in animal tumor models and human brain tumor. Proc. Natl. Acad. Sci. USA 2007, 104, 10152–10157. [Google Scholar] [CrossRef]
  110. Szasz, A.M.; Arrojo, E.E.; Fiorentini, G.; Herold, M.; Herold, Z.; Sarti, D.; Dank, M. Meta-Analysis of Modulated Elec-tro-Hyperthermia and Tumor Treating Fields in the Treatment of Glioblastomas. Cancers 2023, 15, 880. [Google Scholar] [CrossRef] [PubMed]
  111. Stupp, R.; Taillibert, S.; Kanner, A.; Read, W.; Steinberg, D.M.; Lhermitte, B.; Toms, S.; Idbaih, A.; Ahluwalia, M.S.; Fink, K.; et al. Effect of tumor-treating fields plus maintenance temozolomide vs maintenance temozolomide alone on survival in patients with glioblastoma, A randomized clinical trial. JAMA 2017, 318, 2306–2316. [Google Scholar] [CrossRef] [PubMed]
  112. Van Gool, S.W.; Makalowski, J.; Fiore, S.; Sprenger, T.; Prix, L.; Schirrmacher, V.; Stuecker, W. Randomized controlled immunotherapy clinical trials for GBM challenged. Cancers 2021, 13, 32. [Google Scholar] [CrossRef]
  113. Pastore, C.; Fioranelli, M.; Roccia, M.G. Rescue therapy in patient with glioblastoma multiforme combining chemotherapy, hyperthermia, phytotherapy. J. Integr. Oncol. 2017, 6, 199. [Google Scholar]
  114. Hager, E.D.; Birkenmeier, J. Glioblastoma multiforme Grad IV: Regionale Tiefenhyperthermie, Antiangiogenese mit Thalidomid, Hochdosis-Ascorbinsäureinfusionen und komplementäre Therapie. Dtsch. Z. für Onkol. 2006, 38, 133–135. [Google Scholar] [CrossRef]
  115. Mohme, M.; Maire, C.L.; Schliffke, S.; Joosse, S.A.; Alawi, M.; Matschke, J.; Schüller, U.; Dierlamm, J.; Martens, T.; Pantel, K.; et al. Molecular Profiling of an Osseous Metastasis in Glioblastoma During Checkpoint Inhibition: Potential Mechanisms of Immune Escape. Acta Neuropathol. Commun. 2020, 8, 28. [Google Scholar] [CrossRef]
  116. Fiorentini, G.; Sarti, D.; Casadei, V. Modulated electro-hyperthermia (mEHT) [oncothermia®] protocols as complementary treatment. Oncothermia J. 2019, 25, 85–115. [Google Scholar]
  117. Szasz, A.M.; Arkosy, P.; Arrojo, E.E.; Bakacs, T.; Balogh, A.; Barich, A.; Borbenyi, E.; Chi, K.H.; Csoszi, T.; Daniilidis, L.; et al. Guidelines for Local Hyperthermia Treatment in Oncology. In Challenges and Solutions of Oncological Hyperthermia; Szasz, A., Ed.; Cambridge Scholars: Newcastle upon Tyne, UK, 2020; Volume 2, pp. 32–71. Available online: https://www.cambridgescholars.com/challenges-and-solutions-of-oncological-hyperthermia (accessed on 9 October 2020).
  118. Fiorentini, G.; Sarti, D.; Ranieri, G.; Gadaleta, C.D.; Fiorentini, C.; Milandri, C.; Mambrini, A.; Guadagni, S. Modulated electro-hyperthermia in stage III and IV pancreatic cancer: Results of an observational study on 158 patients. World J. Clin. Oncol. 2021, 12, 1064–1071. [Google Scholar] [CrossRef] [PubMed]
  119. Volovat, C.; Volovat, S.R.; Scripcaru, V.; Miron, L. Second-line chemotherapy with gemcitabine and oxaliplatin in combination with loco-regional hyperthermia (EHY-2000) in patients with refractory metastatic pancreatic cancer—Preliminary results of a prospective trial. Rom. Rep. Phys. 2014, 66, 166–174. [Google Scholar]
  120. Dani, A.; Varkonyi, A.; Magyar, T.; Szasz, A. Clinical study for advanced pancreas cancer treated by oncothermia. Forum Hyperthermie 2008, 1, 13–20. [Google Scholar]
  121. Douwes, F.R. Thermochemotherapy of the advanced pancreas carcinoma. Biol. Med. 2006, 35, 126–130. [Google Scholar]
  122. Son, B.; Jeon, J.; Lee, S.; Kim, H.; Kang, H.; Youn, H.; Jo, S.; Youn, B. Radiotherapy in combination with hyperthermia suppresses lung cancer progression via increased NR4A3 and KLF11 expression. Int. J. Radiat Biol. 2019, 95, 1696–1707. [Google Scholar] [CrossRef] [PubMed]
  123. Yeo, S.-G. Definitive radiotherapy with concurrent oncothermia for stage IIIB non-small-cell lung cancer: A case report. Exp. Ther. Med. 2015, 10, 769–772. [Google Scholar] [CrossRef] [PubMed]
  124. Kim, Y.-P.; Choi, Y.; Kim, S.; Park, Y.-S.; Oh, I.-J.; Kim, K.-S.; Kim, Y.-C. Conventional cancer treatment alone or with regional hyperthermia for pain relief in lung cancer: A case-control study. Complement. Ther. Med. 2015, 23, 381–387. [Google Scholar] [CrossRef]
  125. Lee, D.-Y.; Park, S.-J.; Jung, H.-C.; Byun, E.S.; Haam, S.J.; Lee, S.S. The Outcome of the Chemotherapy and Oncothermia for Far Advanced Adenocarcinoma of the Lung: Case reports of four patients. Adv. Lung Cancer 2015, 4, 1–7. [Google Scholar] [CrossRef]
  126. Lee, D.Y. Complete Remission of SCLC with Chemotherapy and Oncothermia (Case report). Oncothermia J. 2012, 5, 43–51. [Google Scholar]
  127. Szasz, A. Current status of oncothermia therapy for lung cancer. Korean J. Thorac. Cardiovasc. Surg. 2014, 47, 77–93. [Google Scholar] [CrossRef]
  128. Roussakow, S.V. Systematic Review of Brain Glioma and Lung Cancer Trials with Modulated Electro-Hyperthermia, with Meta-Analysis and Economic Evaluation (Level II Evidence). Oncothermia J. 2017, 20, 170–216. [Google Scholar]
  129. Dani, A.; Varkonyi, A.; Magyar, T.; Szasz, A. Clinical study for advanced non-small-cell lung-cancer treated by oncothermia. Oncothermia J. 2009, 3, 40–49. [Google Scholar]
  130. Roussakow, S.V. Pharmacoeconomic study of oncothermia (modulated electro-hyperthermia) in the treatment of lung cancer. Oncothermia J. 2016, 18, 116–138. [Google Scholar]
  131. Dani, A.; Varkonyi, A.; Magyar, T.; Kalden, M.; Szasz, A. Clinical study for advanced pancreas cancer treated by oncothermia. Oncothermia J. 2012, 6, 11–25. [Google Scholar]
  132. Hager, E.D.; Dziambor, H.; Höhmann, D.; Gallenbeck, D.; Stephan, M.; Popa, C. Deep hyperthermia with radiofrequencies in patients with liver metastases from colorectal cancer. Anticancer Res. 1999, 19, 3403–3408. [Google Scholar] [PubMed]
  133. Szasz, A.; Szasz, N.; Szasz, O. Oncothermia—Principles and Practices; Springer Science: Heidelberg, Germany, 2010. [Google Scholar]
  134. Minnaar, C.A.; Kotzen, J.A.; Naidoo, T.; Tunmer, M.; Sharma, V.; Vangu, M.-D.; Baeyens, A. Analysis of the effects of mEHT on the treatment-related toxicity and quality of life of HIV-positive cervical cancer patients. Int. J. Hyperth. 2020, 37, 263–272. [Google Scholar] [CrossRef]
  135. Falk, R.E.; Moffat, F.L.; Lawler, M.; Heine, J.; Makowka, L.; Falk, J.A. Combination therapy for resectable and unresectable adenocarcinoma of Pancreas. Cancer 1986, 57, 685–688. [Google Scholar] [CrossRef]
  136. Andocs, G.; Szasz, A.; Szasz, I.; Szasz, N. Tumor Vaccination Patent, 6 October 2020, A1, USA. Available online: http://www.freepatentsonline.com/20150217099.pdf (accessed on 15 April 2021).
  137. Kleef, R. Hyperthermic Oncology. Oncothermia J. 2018, 24, 270–302. [Google Scholar]
  138. Chi, K.-H. Tumour-Directed Immunotherapy: Clinical Results of Radiotherapy with Modulated Electro-Hyperthermia. In Challenges and Solutions of Oncological Hyperthermia; Szasz, A., Ed.; Cambridge Scholars: Newcastle upon Tyne, UK, 2020; Volume 12, pp. 206–226. [Google Scholar]
  139. Pang, C.L.K. The Immune Regulating Effect of Hyperthermia in Combination with TCM on Cancer Patients. Oncothermia J. 2016, 18, 170–179. [Google Scholar]
  140. Minnaar, C.A.; Kotzen, J.A.; Ayeni, O.A.; Vangu, M.-D.-T.; Baeyens, A. Potentiation of the abscopal effect by modulated electro-hyperthermia in locally advanced cervical cancer patients. Front. Oncol. 2020, 10, 376. [Google Scholar] [CrossRef]
  141. Kleef, R.; Moss, R.; Szasz, M.; Bohdjalian, A.; Bojar, H.; Bakacs, T. Complete Clinical Remission of Stage IV Triple-Negative Breast Cancer Lung Metastasis Administering Low-Dose Immune Checkpoint Blockade in Combination With Hyperthermia and Interleukin-2. Integr. Cancer Ther. 2018, 17, 1297–1303. [Google Scholar] [CrossRef] [PubMed]
  142. Kleef, R.; Kekic, S.; Ludwig, N. Successful treatment of advanced ovarian cancer with thermochemotherapy and adjuvant immune therapy. Case Rep. Oncol. 2012, 5, 212–215. [Google Scholar] [CrossRef] [PubMed]
  143. Chi, M.-S.; Wu, J.-H.; Shaw, S.; Wu, C.-J.; Chen, L.-K.; Hsu, H.-C.; Chi, K.-H. Marked local and distant response of heavily treated breast cancer with cardiac metastases treated by combined low dose radiotherapy, low dose immunotherapy and hyperthermia: A case report. Ther. Radiol. Oncol. 2021, 5, 17. [Google Scholar] [CrossRef]
  144. Chi, M.-S.; Mehta, M.P.; Yang, K.-L.; Lai, H.-C.; Lin, Y.-C.; Ko, H.-L.; Wang, Y.-S.; Liao, K.-W.; Chi, K.-H. Putative abscopal effect in three patients treated by combined radiotherapy and modulated electrohyperthermia. Front. Oncol. 2020, 10, 254. [Google Scholar] [CrossRef] [PubMed]
  145. Schirrmacher, V.; Lorenzen, D.; Van Gool, S.W.; Stuecker, W. A new strategy of cancer immunotherapy combining hyperthermia/oncolytic virus pretreatment with specific autologous anti-tumor vaccination—A review. Austin Oncol. Case Rep. 2017, 2, 1006. [Google Scholar]
  146. Schirrmacher, V.; Stücker, W.; Lulei, M.; Bihari, A.S.; Sprenger, T. Long-term survival of a breast cancer patient with extensive liver metastases upon immune and virotherapy: A case report. Immunotherapy 2015, 7, 855–860. [Google Scholar] [CrossRef] [PubMed]
  147. Van Gool, S.; Makalowski, J.; Marko, M. Multimodal immunotherapy for patients with ovarian cancer. Oncothermia J. 2019, 27, 138–152. [Google Scholar]
  148. Van Gool, S.; Makalowski, J.; Feyen, O. Can we monitor immunogenic cell death (ICD) induced with modulated electrohyperthermia and oncolytivc virus injections? Oncothermia J. 2019, 26, 120–125. [Google Scholar]
  149. Van Gool, S.; Makalowski, J.; Marko, M. Hyperthermia as part of multimodal immunotherapy for patients with GBM. Oncothermia J. 2019, 27, 122–137. [Google Scholar]
  150. Van Gool, S.W.; Makalowski, J.; Domogalla, M.P.; Marko, M.; Feyen, O.; Sprenger, K.; Schirrmacher, V.; Stuecker, W. Personalised Medicine in Glioblastoma Multiforme. In Challenges and Solutions of Oncological Hyperthermia; Szasz, A., Ed.; Cambridge Scholars: Newcastle upon Tyne, UK, 2020; Volume 7, pp. 126–158. [Google Scholar]
  151. Van Gool, S.W.; Makalowski, J.; Feyen, O.; Prix, L.; Schirrmacher, V.; Stuecker, W. The induction of immunogenic cell death (ICD) during maintenance chemotherapy and subsequent multimodal immunotherapy for glioblastoma (GBM). Austin Oncol. Case Rep. 2018, 3, 1010. [Google Scholar]
  152. Van Gool, S.W.; Makalowski, J.; Bonner, E.R.; Feyen, O.; Domogalla, M.P.; Prix, L.; Schirrmacher, V.; Nazarian, J.; Stuecker, W. Addition of multimodal immunotherapy to combination treatment strategies for children with DIPG: A single institution experience. Medicines 2020, 7, 29. [Google Scholar] [CrossRef] [PubMed]
  153. Szasz, A. The capacitive coupling modalities for oncological hyperthermia. Open J. Biophys. 2021, 11, 252–313. [Google Scholar] [CrossRef]
  154. Szasz, A. Towards the immunogenic hyperthermic action: Modulated electro-hyperthermia. Clin. Oncol. Res. 2020, 3, 5–6. [Google Scholar]
  155. Vancsik, T.; Kovago, C.; Kiss, E.; Papp, E.; Forika, G.; Benyo, Z.; Meggyeshazi, N.; Krenacs, T. Modulated electro-hyperthermia induced loco-regional and systemic tumor destruction in colorectal cancer allografts. J. Cancer 2018, 9, 41–53. [Google Scholar] [CrossRef]
  156. Tsang, Y.-W.; Huang, C.-C.; Yang, K.-L.; Chi, M.-S.; Chiang, H.-C.; Wang, Y.-S.; Andocs, G.; Szasz, A.; Li, W.-T.; Chi, K.-H. Improving immunological tumor microenvironment using electro-hyperthermia followed by dendritic cell immunotherapy. BMC Cancer 2015, 15, 708. [Google Scholar] [CrossRef]
  157. Qin, W.; Akutsu, Y.; Andocs, G.; Suganami, A.; Hu, X.; Yusup, G.; Komatsu-Akimoto, A.; Hoshino, I.; Hanari, N.; Mori, M.; et al. Modulated electro-hyperthermia enhances dendritic cell therapy through an abscopal effect in mice. Oncol. Rep. 2014, 32, 2373–2379. [Google Scholar] [CrossRef]
  158. Minnaar, C. Challenges Associated with Hyperthermia. In Challenges and Solutions of Oncological Hyperthermia; Szasz, A., Ed.; Cambridge Scholars: Newcastle upon Tyne, UK, 2020; Volume 1, pp. 1–31. [Google Scholar]
  159. Datta, K. Application of SWOT-TOWS matrix and analytical hierarchy process (AHP) in the Formulation of geoconservation and geotourism development strategies for Mama Bhagne Pahar: An important geomorphosite in West Bengal, India. Geoheritage 2020, 12, 45. [Google Scholar] [CrossRef]
  160. Dutz, S.; Hergt, R. Magnetic nanoparticle heating and heat transfer on a microscale: Basic principles, realities and physical limitations of hyperthermia for tumour therapy. Int. J. Hyperth. 2013, 29, 790–800. [Google Scholar] [CrossRef]
  161. Singh, M.; Ma, R.; Zhu, L. Quantitative evaluation of effects of coupled temperature elevation, thermal damage, and enlarged porosity on nanoparticle migration in tumors during magnetic nanoparticle hyperthermia. Int. Commun. Heat Mass Transf. 2021, 126, 105393. [Google Scholar] [CrossRef]
  162. Szasz, O.; Szigeti, G.P.; Szasz, A.M. Electrokinetics of temperature for development and treatment of effusions. Adv. Biosci. Biotechnol. 2017, 8, 434–449. [Google Scholar] [CrossRef]
  163. Szasz, O.; Szigeti, G.P.; Szasz, A.; Benyo, Z. Role of electrical forces in angiogenesis. Op. J. Biophys. 2018, 8, 49–67. [Google Scholar] [CrossRef]
  164. Chen, C.-C.; Chen, C.-L.; Li, J.-J.; Chen, Y.-Y.; Wang, C.-Y.; Wang, Y.-S.; Chi, K.-H.; Wang, H.-E. The presence of gold nanoparticles in cells associated with the cell-killing effect of modulated electro-hyperthermia. ACS Appl. Bio Mater. 2019, 2, 3573–3581. [Google Scholar] [CrossRef]
  165. Singh, M. Biological heat and mass transport mechanisms behind nanoparticles migration revealed under microCT image guidance. Int. J. Therm. Sci. 2023, 184, 107996. [Google Scholar] [CrossRef]
  166. Zotin, A.I. Thermodynamic Bases of Biological Processes; Cambridge University Press: Cambridge, UK, 2010. [Google Scholar]
  167. Vaupel, P.; Kallinowski, F.; Okunieff, P. Blood flow, oxygen and nutrient supply, and microenvironment of human tumors: A review. Cancer Res. 1989, 49, 6449–6465. [Google Scholar] [PubMed]
  168. Dudar, T.E.; Jain, R.K. Differential response of normal and tumor microcirculation to hyperthermia. Cancer Res. 1984, 44, 605–612. [Google Scholar]
  169. Song, C.W.; Lokshina, A.; Rhee, J.G.; Patten, M.; Levitt, S.H. Implication of blood-flow in hyperthermic treatment of tumors. IEEE Trans. Biomed. Eng. 1984, 31, 9–16. [Google Scholar] [CrossRef]
  170. Qwaider, Y.Z.; Sell, N.M.; Stafford, C.E.; Kunitake, H.; Cusack, J.C.; Ricciardi, R.; Bordeianou, L.G.; Deshpande, V.; Goldstone, R.N.; Cauley, C.E.; et al. Infiltrating Tumor Border Configuration is a Poor Prognostic Factor in Stage II and III Colon Adenocarcinoma. Ann. Surg. Oncol. 2021, 28, 3408–3414. [Google Scholar] [CrossRef]
  171. Sing, M. Incorporating vascular-stasis based blood perfusion to evaluate the thermal signatures of cell-death using modified Arrhenius equation with regeneration of living tissues during nanoparticle-assisted thermal therapy. Int. Commun. Heat Mass Transf. 2022, 135, 106046. [Google Scholar] [CrossRef]
  172. Vaupel, P.W.; Kelleher, D.K. Metabolic Status and Reaction to Heat of Normal and Tumor Tissue. In Thermo-Radiotherapy and Thermo-Chemotherapy. Biology, Physiology and Physics; Seegenschmiedt, M.-H., Fessenden, P., Vernon, C.C., Eds.; Springer: Berlin/Heidelberg, Germany, 1996; Volume 1, pp. 157–176. [Google Scholar]
  173. Nomura, D.K.; Long, J.Z.; Niessen, S.; Hoover, H.S.; Ng, S.-W.; Cravatt, B.F. Monoacylglycerol lipase regulates a fatty acid network that promotes cancer pathogenesis. Cell 2010, 140, 49–61. [Google Scholar] [CrossRef]
  174. Warburg, O. Oxygen, The Creator of Differentiation, Biochemical Energetics. In The Prime Cause and Prevention of Cancer: Revised Lecture at the Meeting of the Nobel-Laureates on 30 June 1966, Lindau, Lake Constance, Germany; Academic Press: New York, NY, USA, 1996. [Google Scholar]
  175. Peppicelli, S.; Andreucci, E.; Ruzzolini, J.; Bianchini, F.; Calorini, L. FDG uptake in cancer: A counting debate. Theranostics 2020, 10, 2944–2948. [Google Scholar] [CrossRef]
  176. Kabakov, A.E.; Gabai, V.L. Heat Shock Proteins and Cytoprotection: ATP-Deprived Mammalian Cells. (Molecular Biology Intelligence Unit); Springer: Berlin, Heidelberg, Germany, 1997. [Google Scholar]
  177. Dikomey, E.; Franzke, J. Effect of heat on induction and repair of DNA strand breaks in X-irradiated CHO cells. Int. J. Radiat. Biol. 1992, 61, 221–233. [Google Scholar] [CrossRef] [PubMed]
  178. Oei, A.L.; Vriend, L.E.M.; Crezee, J. Effects of hyperthermia on DNA repair pathways: One treatment to inhibit them all. Radiat. Oncol. 2015, 10, 165. [Google Scholar] [CrossRef]
  179. Okumura, Y.; Ihara, M.; Shimasaki, T.; Takeshita, S.; Okaichi, K. Heat Inactivation of DNA-Dependent Protein Kinase: Possible Mechanism of Hyperthermic Radio-Sensitization. In Thermotherapy for Neoplasia, Inflammation, and Pain; Kosaka, M., Sugahara, T., Schmidt, K.L., Simon, E., Eds.; Springer: Tokyo, Japan, 2001; pp. 420–423. [Google Scholar]
  180. Hayashi, S.; Kano, E.; Hatashita, M.; Ohtsubo, T.; Katayama, K.; Matsumoto, H. Fundamental Aspects of Hyperthermia on Cellular and Molecular Levels; Kosaka, M., Sugahara, T., Schmidt, K.L., Simon, E., Eds.; Springer: Tokyo, Japan, 2001; pp. 335–345. [Google Scholar]
  181. Kühl, N.M.; Rensing, L. Heat shock effects on cell cycle progression. Cell. Mol. Life Sci. CMLS 2000, 57, 450–463. [Google Scholar] [CrossRef] [PubMed]
  182. Gascoyne, P.R.C.; Pethig, R.; Szent-Gyorgyi, A. Water structure-dependent charge transport in proteins. Proc. Natl. Acad. Sci. USA 1981, 78, 261–265. [Google Scholar] [CrossRef] [PubMed]
  183. Kotnik, T.; Miklavcic, D. Theoretical evaluation of the distributed power dissipation in biological cells exposed to electric fields. Bioelectromagnetics 2000, 21, 385–394. [Google Scholar] [CrossRef]
  184. Gombos, I.; Crul, T.; Piotto, S.; Güngör, B.; Török, Z.; Balogh, G.; Péter, M.; Slotte, J.P.; Campana, F.; Pilbat, A.-M.; et al. Membrane-Lipid therapy in operation: The HSP co-inducer GBP-15 activates stress signal transduction pathways by remodeling plasma membrane rafts. PLoS ONE 2011, 6, e28818. [Google Scholar] [CrossRef]
  185. Beachy, S.H.; Repasky, E.A. Toward establishment of temperature thresholds for immunological impact of heat exposure in humans. Int. J. Hyperth. 2011, 27, 344–352. [Google Scholar] [CrossRef]
  186. Tell, R.A.; Tell, C.A. Perspectives on setting limits for RF contact currents: A commentary. Tell Tell Biomed. Eng. OnLine 2018, 17, 2. [Google Scholar] [CrossRef]
Figure 1. Selective and theranostic behaviour of mEHT. (A) The malignancy is localized and proven with conventional methods. (B) The RF current macroscopically selects the tumor by electric conductivity and microscopically by dielectric permittivity. The transmembrane proteins thermally and nonthermally absorb the energy in the malignant cells. The pathologic irregularities further increase the selection by applying RF current. (C) The nanoscopic membrane rafts locally heat the malignant cells, and the optimized modulated RF current makes the desired molecular changes for DAMP and ICD. The DAMP molecules induce antigen presentation and, as a consequence, antitumoral killer and helper T-cells appear, working as a tumor-specific vaccination.
Figure 1. Selective and theranostic behaviour of mEHT. (A) The malignancy is localized and proven with conventional methods. (B) The RF current macroscopically selects the tumor by electric conductivity and microscopically by dielectric permittivity. The transmembrane proteins thermally and nonthermally absorb the energy in the malignant cells. The pathologic irregularities further increase the selection by applying RF current. (C) The nanoscopic membrane rafts locally heat the malignant cells, and the optimized modulated RF current makes the desired molecular changes for DAMP and ICD. The DAMP molecules induce antigen presentation and, as a consequence, antitumoral killer and helper T-cells appear, working as a tumor-specific vaccination.
Cancers 15 04569 g001
Figure 2. The protocol for the GBM treatment. The modulation of the carrier frequency, and the energy load had to be adapted by the patient (green columns).
Figure 2. The protocol for the GBM treatment. The modulation of the carrier frequency, and the energy load had to be adapted by the patient (green columns).
Cancers 15 04569 g002
Figure 3. The survival results at different times and various clinics. The Kaplan–Meier probability comparison showed no statistical difference between the different clinical studies, so the data may be pooled containing 325 patients altogether. The average median value was significantly higher than the SEER database (data from [61,99]). (① = [99], ② = [96] ③ = [95] ④ = [101] ⑤ = [103], ⑥ = [97], ⑦ = SEER (NCI, USA) data from [99].).
Figure 3. The survival results at different times and various clinics. The Kaplan–Meier probability comparison showed no statistical difference between the different clinical studies, so the data may be pooled containing 325 patients altogether. The average median value was significantly higher than the SEER database (data from [61,99]). (① = [99], ② = [96] ③ = [95] ④ = [101] ⑤ = [103], ⑥ = [97], ⑦ = SEER (NCI, USA) data from [99].).
Cancers 15 04569 g003
Figure 4. Comparison of mEHT pooled GBM survival probability with the literature [105]. ① is the reference arm (radiotherapy (RT) alone) and ② is the active arm, RT + TMZ.
Figure 4. Comparison of mEHT pooled GBM survival probability with the literature [105]. ① is the reference arm (radiotherapy (RT) alone) and ② is the active arm, RT + TMZ.
Cancers 15 04569 g004
Figure 5. The invasive brachytherapy and invasive hyperthermia (iHT) in brain for GBM [106] do not differ from the mEHT pooled data. ① = brachytherapy (bTh) alone and ② = bTh + iHT.
Figure 5. The invasive brachytherapy and invasive hyperthermia (iHT) in brain for GBM [106] do not differ from the mEHT pooled data. ① = brachytherapy (bTh) alone and ② = bTh + iHT.
Cancers 15 04569 g005
Figure 6. Comparison of the pooled data of six mEHT studies to TTF+TMZ survival data [72,111], where ① shows the result of the active study arm TTF+TMZ and ② is the reference with TMZ alone.
Figure 6. Comparison of the pooled data of six mEHT studies to TTF+TMZ survival data [72,111], where ① shows the result of the active study arm TTF+TMZ and ② is the reference with TMZ alone.
Cancers 15 04569 g006
Figure 7. Comparison of the survival studies of advanced pancreas carcinomas ① [75] (n = 106), ② [99] (n = 25), ③ [120] (n = 99), ④ [76] (n = 78), ⑤ [120] (n = 34), ⑥ SEER (NCI, USA) data from [99].
Figure 7. Comparison of the survival studies of advanced pancreas carcinomas ① [75] (n = 106), ② [99] (n = 25), ③ [120] (n = 99), ④ [76] (n = 78), ⑤ [120] (n = 34), ⑥ SEER (NCI, USA) data from [99].
Cancers 15 04569 g007
Figure 8. Comparison of the NSCLC survivals from large databases [92,130]. ① mEHT pool, ② Best RT+ CTx, ③ SEER (NCI, USA), ④ Eurocare-5, ⑤ East-EU, Eurocare-5.
Figure 8. Comparison of the NSCLC survivals from large databases [92,130]. ① mEHT pool, ② Best RT+ CTx, ③ SEER (NCI, USA), ④ Eurocare-5, ⑤ East-EU, Eurocare-5.
Cancers 15 04569 g008
Figure 9. The measured NSCLC survival probabilities. ① = [127], ② = [129], ③ = [131], ④ = [61,99], ⑤ = SEER (NCI, USA) data from [99].
Figure 9. The measured NSCLC survival probabilities. ① = [127], ② = [129], ③ = [131], ④ = [61,99], ⑤ = SEER (NCI, USA) data from [99].
Cancers 15 04569 g009
Figure 10. The colorectal cancer survival plots. ① = [99], ② = [133], ③ = [132], ④ = [132], ⑤ = [82].
Figure 10. The colorectal cancer survival plots. ① = [99], ② = [133], ③ = [132], ④ = [132], ⑤ = [82].
Cancers 15 04569 g010
Figure 11. Heating heterogeneity. The characteristic sizes and the expected temperatures are shown in the various steps.
Figure 11. Heating heterogeneity. The characteristic sizes and the expected temperatures are shown in the various steps.
Cancers 15 04569 g011
Table 1. The dose escalation of the GBM treatment with mEHT. Advanced GBM 3rd and 4th line treatment was studied. The complementary chemotherapy was nitrosourea drug Nimustine (ACNU) 90   m g / m 2 . Patients were grouped in 4 disjunct study arms by dosing for 6 weeks cycles.
Table 1. The dose escalation of the GBM treatment with mEHT. Advanced GBM 3rd and 4th line treatment was studied. The complementary chemotherapy was nitrosourea drug Nimustine (ACNU) 90   m g / m 2 . Patients were grouped in 4 disjunct study arms by dosing for 6 weeks cycles.
GroupNumber of PatientsmEHT/Week (6 Cycles)
142
243
344
435
Table 2. Survival results of the Phase III uterus/cervix cancer study [67,68].
Table 2. Survival results of the Phase III uterus/cervix cancer study [67,68].
Groups Number of Patients% Average Age (y)3 y Overall Survival (%)p-Value
All 210100
RT + ChT aloneHIV positive5552.950.633.70.04
HiV negative4947.1
RT + ChT + mEHTHIV positive5249.149.244
HiV negative5450.9
Table 3. Results of retrospective double-arm observational study for cervix carcinoma. (CR-complete remission, DFS disease-free survival, NED—no evidence of disease.
Table 3. Results of retrospective double-arm observational study for cervix carcinoma. (CR-complete remission, DFS disease-free survival, NED—no evidence of disease.
Groups Number of Patients% OS after 5 yCR with Lymph Node Mets.CR (NED) DFS after 5 yDFS with Lymph Node Mets. after 5 y
All 95100%p-value%p-value%p-value%p-value%p-value
RT + ChT alone505379.50.079450.0377580.0315730.166730.166
RT + ChT + mEHT40428171828080
Table 4. The response data of the treatments. AST—astrocytoma (Grade III), CR—complete remission, PR—partial remission, SD—stable disease, PD—progressive disease, OS—overall survival.
Table 4. The response data of the treatments. AST—astrocytoma (Grade III), CR—complete remission, PR—partial remission, SD—stable disease, PD—progressive disease, OS—overall survival.
Response Astrocytoma Glioblastoma
without mEHT (n, %)with mEHT (n, %)without mEHT (n, %)with mEHT (n, %)
n%n%n%n%
CR628.626.922.413.4
PR14.81034.522.4620.7
SD523.8931.01315.31137.9
PD838.1620.76374.11137.9
No data14.826.955.900.0
OS median (months)17721215
OS range3–1203–1562–842–108
p-value0.00060.026
Table 5. The main parameters of two survival studies [77,78] (*—macro metastases, ** possible micro metastases.).
Table 5. The main parameters of two survival studies [77,78] (*—macro metastases, ** possible micro metastases.).
Response Fiorentini et al. Petenyi et al.
without mEHT (n, %)with mEHT (n, %)without mEHT (n, %)with mEHT (n, %)
n%n%n%n%
Patients no.67393939
Males3856.72461.51946.21846.2
Females2943.31538.52053.82155.8
Age (mean, y)6661.866.0265.9
Distant metastasis *3755.22564.12461.52051.3
Non metastatic **3044.81435.91538.51948.7
Gemcitabine combination6495.52769.23179.53179.5
Other complementary34.51230.8820.5820.5
OS median (months)10.91810.5817.02
OS range0.4–55.41.5–682.4–48.84.4–47.1
p0.001650.0301
Table 6. The GBM studies which were used for KM comparison in comparison Figure 1.
Table 6. The GBM studies which were used for KM comparison in comparison Figure 1.
No.Number of PatientsTreatmentsOS Median (Months)Reference
135mEHT + RT + ChT + BST26.4Parmar, et al. 2020 [99]
228mEHT + RT + ChT + BSC, (palliative)14Fiorentini, et al. 2018 [96]
392mEHT + RT + ChT20.4Sahinbas, et al. 2007 [95]
4126mEHT + RT + ChT20.3Hager, et al. 2008 [101]
529mEHT + RT + ChT14Szasz, et al. 2010 [103]
619ChT (ACNU)21.8Douwes, et al. 2006 [97]
Table 7. The pancreas carcinoma studies which are used for KM comparison in comparison in Figure 5.
Table 7. The pancreas carcinoma studies which are used for KM comparison in comparison in Figure 5.
No.Number of PatientsTreatmentsOS Median (Months)Reference
139GMZ combination with mEHT18Fiorentini et al. 2019, [75]
227GMZ combination with mEHT13.2Parmar et al. 2020 [99]
399GMZ combination with mEHT12Dani et al. 2012 [120]
439GMZ combination with mEHT17Petenyi et al. 2021 [76]
534GMZ combination without mEHT6.5Dani et al. 2012 [120]
Table 8. The survival data of NSCLC studies.
Table 8. The survival data of NSCLC studies.
No.Number of PatientsTreatmentsOS Median (Months)Reference
175RT + ChT + OP with mEHT16.4Szasz, 2014 [127]
2197RT + ChT + OP with mEHT15.6Dani, et al. 2012 [129]
361RT + ChT + OP with mEHT16.4Dani, et al. 2009 [131]
454RT + ChT + BSC with mEHT18Parmar, et al. 2020 [99]
Table 9. Survival data of colorectal studies.
Table 9. Survival data of colorectal studies.
No.Number of PatientsTreatmentsOS Median (Months)Reference
179ChT with mEHT48Parmar, et al. 2020 [99]
2218OP + ChT with mEHT28.5Szasz, et al. 2010 [133]
350BSC with mEHT25Hager, et al. 2020 [132]
430ChT + BSC with mEHT23Hager, et al. 2020 [132]
540ChT with mEHT21.4Ranieri, et al. 2020 [82]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lee, S.-Y.; Lorant, G.; Grand, L.; Szasz, A.M. The Clinical Validation of Modulated Electro-Hyperthermia (mEHT). Cancers 2023, 15, 4569. https://doi.org/10.3390/cancers15184569

AMA Style

Lee S-Y, Lorant G, Grand L, Szasz AM. The Clinical Validation of Modulated Electro-Hyperthermia (mEHT). Cancers. 2023; 15(18):4569. https://doi.org/10.3390/cancers15184569

Chicago/Turabian Style

Lee, Sun-Young, Gergo Lorant, Laszlo Grand, and Attila Marcell Szasz. 2023. "The Clinical Validation of Modulated Electro-Hyperthermia (mEHT)" Cancers 15, no. 18: 4569. https://doi.org/10.3390/cancers15184569

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop